You are on page 1of 243

AERODYNAMIC PREDICTIVE METHODS

AND THEIR VALIDATION IN


HYPERSONIC FLOWS
AK SREEKANTH

AERODYNAMIC PREDICTIVE METHODS


AND THEIR VALIDATION IN
HYPERSONIC FLOWS
AK SREEKANTH

Defence Research & Development Organisation


Ministry of Defence
New Delhi 110 011
2003

DRDO MONOGRAPH SERIES


AERODYNAMIC PREDICTIVE METHODS AND THEIR
VALIDATION IN HYPERSONIC FLOWS
AK SREEKANTH
Series Editors
Editor-in-Chief
Dr Mohinder Singh

Editors
Dr JP Singh, A Saravanan

Coordinator
Ashok Kumar

Cover Design
A Saravanan

Asst. Editor
Ramesh Chander

Editorial Asst.
AK Sen, Kumar Amar Nath

Production
Printing
JV Ramakrishna, SK Tyagi

Marketing
RK Dua, Rajpal Singh

Cataloguing in Publication
SREEKANTH, A.K.
Aerodynamic predictive methods and their validation in hypersonic
flows.
DRDO monograph series.
Includes index and bibliography.
ISBN 81-86514-11-2
1. Aerodynamics 2. Hypersonic flows I. Title (Series)
629.132.306.072

2003, Defence Scientific Information & Documentation Centre (DESIDOC),


Defence R&D Organisation, Delhi-110 054.
All rights reserved. Except as permitted under the Indian Copyright Act 1957, no
part of this publication may be reproduced, distributed or transmitted, stored in a
database or a retrieval system, in any form or by any means, electronic, mechanical,
photocopying, recording, or otherwise, without the prior written permission of the
publisher.
The views expressed in the book are those of the author only. The editors or publisher
do not assume responsibility for the statements/opinions expressed by the author.

Printed and published by Director, DESIDOC, Metcalfe House, Delhi-110 054.

CONTENTS
Preface

xi

Acknowledgement

PART - I

AERODYNAMIC PREDICTIVE METHODS IN


HYPERSONIC FLOWS

xiii

CHAPTER 1
AERODYNAMIC PREDICTIVE METHODS IN HYPERSONIC
FLOWS

CHAPTER 2
METHODS

2.1

Introduction

2.2

Newtonian Theory

2.3

Modified Newtonian Theory

2.4

Embedded Newtonian Flow

2.5

Newtonian & Prandtl-Meyer Mode1

10

2.6

Tangent Wedge & Tangent Cones

13

2.7

Tangent Wedge, Tangent Cone & Delta


Wing Empirical Method

14

2.8

OSU Blunt Body Method

17

2.9

Hankey Flat Surface Empirical Method

17

2.10

Dahlem-Buck Empirical Method

17

2.11

Blast Wave Pressure Increments

18

2.12

Shock Expansion Theory

18

2.12.1

First Order Theory

18

2.12.2

Second Order Shock Expansion Theory (SOSET)

19

2.13

Blunt Bodies of Revolution at Small Angles of Attack

21

2.14

Van Dyke Unified Theory

24

2.15

2-D Airfoil Theory in Hypersonic Flows

25

2.16

High Mach Number Base Pressure

30

References

31

(vi)

CHAPTER 3
AERODYNAMIC CHARACTERISTICS OF VEHICLE
COMPONENTS

33

3.1

Introduction

33

3.2

Body-Alone Aerodynamics

34

3.2.1

Forces & Moments on the Body

34

3.2.2

Axial Force

37

3.2.3

CA

37

3.2.4

C A Base Pressure Coefficient


b

39

3.2.5

Determination of CA N , the Axial Pressure


Coefficient of Nose Portion of the Body

40

3.2.5.1

Pointed Cone

40

3.2.5.2

Pointed Ogive

41

3.2.5.3

Hemispherical Nose

41

3.2.6

Normal Force

41

3.2.6.1

Pointed Cone

41

3.2.6.2

Pointed Ogive

42

3.2.6.3

Hemispherical Nose

42

3.2.6.4

Cylinder

42

3.3

Allen & Perkins Viscous Cross Flow Theory

43

3.4

Moments

43

3.4.1

Pointed Cone

44

3.4.2

Pointed Ogive

44

3.4.3

Hemisphere

45

3.4.4

Circular Cylinder

45

3.5

Wing Alone Aerodynamics

45

3.5.1

Hexagonal Shape Wing Section

45

Skin Friction Coefficient

3.5.1.1

Axial Force

47

3.5.1.2

Normal Force

50

3.5.1.3

Axial Component of the Rudder

51

3.5.1.4

Normal Component (Wing or Rudder)

51

3.5.1.5

Pitching Moment

52

3.5.2

Other Wing Sections

52

3.5.2.1

Airfoil Characteristics by 2-Dimensional


Hypersonic Airfoil Theory

52

References

59

(vii)

CHAPTER 4
SKIN FRICTION FORCE CALCULATION

61

4.1

Introduction

61

4.2

Sommer & Short Method

62

4.3

Van Driest-II Method

63

4.4

Spalding & Chi Method

64

4.5

Empirical Equations

65

References

66

CHAPTER 5
AERODYNAMIC HEATING AT HYPERSONIC SPEEDS

67

5.1

Introduction

67

5.2

Heating Analysis

67

5.3

Stagnation Point Heat Transfer

69

5.3.1

Spherical Nose

69

5.3.2

Cylinder Normal to the Stream

72

5.3.3

Swept Wing Stagnation Line Heat Transfer

72

5.3.4

Perfect Gas

73

5.3.5

Real Gas

74

5.3.6

Heat Transfer Coefficient h

75

5.3.7

Heat Transfer on Flat Surfaces and Fuselage Panels

77

5.4

Heat Transfer Analysis by the Method of Quinn & Gong

80

5.4.1

Stagnation Point Heating Rate

80

5.4.2

Convective Heating Equation for Small or Zero


Pressure Gradient Surfaces

83

5.4.3

Boundary Layer Transition

86

5.5

High Speed Convective Heat Transfer Methodology of


Tauber

87

5.5.1

Stagnation Point Heat Transfer

88

5.5.2

Swept Infinite Cylinder

88

5.5.3

Cone & Flat Plate Heating Rate

89

5.5.3.1

Laminar Boundary Layer

89

5.6

Empirical Equation for Convective Heat Transfer

90

5.6.1

Stagnation Point

91

5.6.2

Flat Plate in Laminar Flow

91

5.6.3

Flat Plate in Turbulent Flow

91

References

92

(viii)

PART - II VALIDATION OF PREDICTION METHODS

93

CHAPTER 6
VALIDATION OF PREDICTION METHODS

95

6.1

North American X-15 Research Aircraft

110

6.1.1

Walker & Wolowiczs Work

115

6.1.2

Lift Characteristics

115

6.1.3

Wing

116

6.1.4

Horizontal Tail

118

6.1.5

Fuselage

118

6.1.6

Pitching-Moment Characteristics

119

6.1.7

Wing & Horizontal Tail

119

6.1.8

Fuselage

128

6.1.9

Maughmer et al. Analysis of X-15

128

6.2

Hypersonic Research Airplane

139

6.3

Space Shuttle Orbiter

156

6.4

Conclusions

158

References

169

PART - III AERODYNAMICS OF RAREFIED GASES

173

CHAPTER 7
AERODYNAMICS OF RAREFIED GASES

175

7.1

Introduction

175

7.2

Free Molecule Flow Analysis

177

7.2.1

Surface Interaction Parameters

177

7.2.2

Forces on an Surface Element in Free Molecule Flow

179

7.3

Aerodynamic Forces for Typical Bodies

187

7.3.1

Flat Plate

187

7.3.2

Infinite Right Circular Cylinder at an Angle of Attack, 191

7.3.3

Sphere

194

7.3.4

Cone Frustrum

195

7.3.5

Spherical Segment

198

7.4

Aerodynamic Forces in Slip & Transitional Flows

200

7.5

Energy Transfer in Free Molecule Flow

203

7.5.1

Equilibrium Temperatures for Simple Shapes

207

(ix)

7.5.2

Heat Transfer for Typical Bodies in Free Molecule Flow 208

7.5.3

Heat Transfer in Slip & Transitional Flow Regimes

212

References

213

Appendix

215

Index

225

PREFACE
This monograph presents a summary of engineering
methods most commonly employed for preliminary aerodynamic
analysis of bodies travelling at hypersonic speeds. To the extent
possible, an attempt has been made to make the present work
self-sufficient. However, references are cited if one is interested
in the source or more details.
The work is in three parts. Part 1 deals with Predictive
Methodology, Part 2 covers Validation of Prediction Methods and
Part 3 the Aerodynamics of Rarefied Gases.
Secunderabad
Date: June 2003

AK Sreekanth

ACKNOWLEDGEMENT
The writing of this monograph has been made possible by
the financial assistance received from the Defence Scientific
Information and Documentation Centre (DESIDOC), Ministry of
Defence, Government of India, New Delhi.
The author would like to place on record his sincere
thanks and appreciation to the following persons.
1.

2.

3.

Prof. M. Maughmer, Department of Aerospace Engineering,


The Pennsylvania State University, University Park, PA.
U.S.A. for permission to freely use the figures and material
from the thesis of his student L.P.Ozoroski and from the
NASP Contractor Report 1104.
Dr. J.Agrell, Head of Experimental Aerodynamics
Department, FFA, Sweden, for permission to include
material in the monograph from the FFA Technical Note AU1661.
Mr. Dan Pappas, Chief Librarian, NASA Ames Research
Center, Moffett Field, CA. for allowing me to use the Ames
Library freely.

PART - I
AERODYNAMIC PREDICTIVE METHODS
IN
HYPERSONIC FLOWS

CHAPTER 1
AERODYNAMIC PREDICTIVE METHODS IN
HYPERSONIC FLOWS
1.1

INTRODUCTION

The conceptual design of an efficient hypersonic cruise


vehicle or a missile requires a detailed knowledge of how various
geometrical configuration parameters affect the aerodynamic
performance of such a vehicle. Besides, it is desirable to have the
ability to compare one configurations performance with another
in a relatively short amount of time. During the preliminary design
phase involved in arriving at feasible configurations for a specified
mission, simple engineering-type empirical and semi-empirical
methods are invariably employed. The expensive and timeconsuming wind tunnel tests and sophisticated computational
techniques are reserved for possible designs evolved from the
preliminary analysis.
A variety of engineering methods applicable to flows at
hypersonic Mach numbers have been reported over the years in
open literature. Each of these methods works well on very specific
types of components. Therefore, it is necessary to choose a
combination of these methods to analyse the complete vehicle
made up of various components, such as body, lifting, and control
surfaces.
The present work is a compilation of some of the wellknown prediction methods, their applicability and limitations.
Examples of the application of a few of these methods to calculate
aerodynamic parameters of some specific components of vehicle
configurations have been made and the results presented. Some
published work on the aerodynamic characteristics of a few of
hypersonic configurations, their predictions and comparison with

Aerodynamic Predictive Methods In Hypersonic Flows

experimental data are discussed in Part II(Chapter 6) of this


monograph, to illustrate the applicability and validity of the
approximate methodology.

2.1

INTRODUCTION

A majority of the methods for calculating the pressure


forces in hypersonic flow are based on non-interfering constant
pressure finite element analysis. The geometry of the
configuration is represented by a system of quadrilateral panels.
The only parameter required to calculate the pressure is the
impact angle of the free stream flow with the panel or the change
in impact angle from one panel to another. The surface elements
that see the oncoming flow directly, are said to be in the impact
region and the others, either shielded by the front portion or other
surrounding elements, are in the shadow region. Depending upon
whether the element is in the impact or shadow region, the
appropriate method is chosen for the analysis. Body components
are typically broken into separate analysis regions. The forward
most body component may have a nose and body region. The rear
most body consists of a body region and probably a blunt base. In
each of these separate regions, an analysis method must be
chosen for the impact flow region and a suitable one for the shadow
region. Similar division is also done for the lifting and control
surfaces. viz., a leading edge region, a surface (mostly flat) region
and a blunt region if the trailing edge is blunt. Most commonly
used methods are listed in Table 1.1.

Table 1.1. List of most commonly used methods


Imwact Flow

Shadow Flow

Newtonian and Modified Newtonian


Embedded Newtonian
Modified Newtonian + Prandtl-Meyer
Tangent wedge and Tangent cone
Tangent wedge-Tangent cone
(empirical)

Newtonian (Cp=O)
Modified Newtonian + Prandtl-Meyer
Prandtl-Meyer from free stream
OSU blunt body empirical '
Van Dyke unified
Contd...

Aerodynamic Predictive Methods and their Validation i n Hypersonic Flows


Imwact Flow
OSU blunt body empirical
Van Dyke unified
2-Dim. Hypersonic airfoil theory

Shadow Flow
High Mach number base pressure
Shock expansion
Rarefied gas flow

Shock-expansion
Input pressure coefficient
Hankey flat surface empirical
Delta wing empirical
Dahlem-Buck empirical
Blast wave
Rarefied gas flow

Brief reviews of some of the above listed methods are discussed


in following pages.

2.2

NEWTONIAN THEORY
Newtonian theory is a local surface inclination method. In
this, the pressure coefficient depends only on the local surface
deflection angle and not on any other aspect of the surrounding
flow field.
Newton originally assumed that the medium around a body
was composed of identical non-interfering independent particles.
When these particles collide with the surface they lose their
normal component of the momentum resulting in a pressure force
on it. After collision, the particles move along the surface with
their tangential component of the momentum unchanged. The
regions of the body that do not see the oncoming particles directly
are said to be in the shadow region and the pressure coefficient in
these regions are normally set equal to zero.

Figure 2.1. Newtonian theory

The normal component of the velocity is V, sin 6 . The


mass of particles striking the surface area A in unit time is
p, V , A sin S . The total normal momentum carried by these

Methods

particles is pm v,2 sin2 6 . This momentum is transferred to the


surface element and acts a s the normal pressure force.

If F / A is interpreted a s the difference in pressure above


the free stream, we have

For blunt bodies in a high Mach number flow, the surface


pressure is fairly well predicted by the above Newtonian theory. It
is to be observed that according to the Newtonian theory the
pressure coefficient is independent of the Mach number, so long a s
the flow is hypersonic.
MODIFIED NEWTONIAN THEORY
In the Modified Newtonian theory (MNT), the pressure
coefficient is written a s

2.3

Various values of K have been suggested depending on


the Mach number, body shape, angle of attack and ratio of
is the
specific heats. The most common one is K = Cpswhere Cps
stagnation pressure coefficient behind a normal shock. For this
case we have

and mainly used for blunt bodies. For


1.839andfor Y = 1.0and M,

-+

1.4 and M, +co, K+

co, K j 2 . 0 .

For pointed cones and ogives, the suggested1 value for K is

where, k = 1 for cones and k = 0 for ogives, d


L,= nose length and a = angle of attack.

body diameter,

Aerodynamic Predictive Methods and their Validationin Hypersonic Flows

For a hemisphere1,

For real gas flows,

for

Y=

1.4, K-2.083,andfor

Y=

1,K=2.0.

Although the Newtonian and Modified Newtonian theories


are mainly applicable for blunt body flows, attempts have been
made to see whether the same form of modified equation could be
made applicable to flat surfaces such a s wings. One such suggested
relation by Hankeyqs
K =1.95 +

0.3925

M : . ~ tan 6

where, 6 is the local flow inclination angle. The above expression


is applicable in the Mach number range of 2 to 22 and angles of
attack from 10" to 90" for surfaces with highly swept leading
edges.
For surface inclinations below 10 degrees, particularly for
wings, Newtonian theory is not applicable. Interaction and induced
pressure effects also become dominant at low angles of attack
requiring a different methodology.
EMBEDDED NEWTONIAN FLOW
For bodies having compression corners on the surfaces
such a s flares or flap controls, the Newtonian theory may not
correctly predict the pressure on these surfaces a s there may be a
oblique shock in front of the ramp if the local flow is supersonic a s
shown in the Fig. 2.2.
Newtonian theory assumes that the bow shock wgve in
front of the body wraps around the body very closely thereby not
giving rise to secondary shock that might be normally present in
local supersonic regions. For configurations like this, a method

2.4

Methods

9
EMBEDDED SHOCK

FLARE

(a)

-EMBEDDED SHOCK

FLAP

(b)
Figure 2.2. Embedded Flows

has been suggested by Sieffj in which the flow over the ramp is
viewed a s an embedded Newtonian impact flow if the flow is not
extensively separated and the ramp sh0c.k wave is thin, with
conditions along the surface of the secondary shock wave a s
initial conditions. According to this postulation, the pressure on
the ramp surface is given by the following expression:
P2 - PI = PI ( ~ sin
1
dl2
where, the subscript 1 refers to conditions along the front surface
of the ramp shock wave, which are the initial conditions for the
application of Newtonian theory on the ramp surface. The above
relation can be expressed in the form of pressure coefficient
based on free stream static and dynamic pressure conditions,

viz.,

41

P2

'P2
4,

Newtonian

10

Aerodynamic Predictive Methods and their Validationin Hypersonic Flows

where, Cp,,,O,i,
is the pressure coefficient given by the usual
Newtonian Impact theory and q the dynamic pressure.
For the application of the above formulation one needs to
know the properties of the stream that is incident on the ramp
surface. Towards this, one can utilize the methods presented in
Sieff, et ~ l .and
, ~ Maslen, et ~ l .or, any
~ other known procedures.

2.5

NEWTONIAN 88 PRANDTL-MEYER MODEL

In this flow model applicable to blunt bodies with a


detached shock, it is assumed that the flow expands around the
body surface to supersonic conditions isentropically from the
stagnation point. Modified Newtonian theory is combined with
Prandtl-Meyer Expansion theory. The technique involves
matching the Modified Newtonian and Prandtl-Meyer Expansion
methods at the point where the pressure gradients calculated
using each method are equal. Downstream of this point the
Prandtl-Meyer Expansion theory is used6. The calculation
procedure is a s follows:
Calculate the ratio of the free stream static to stagnation
pressure behind a normal shock.

According to Newtonian theory

At the stagnation point 6 = 90"

Substitutingfor

YM,~K

from Eqn. 2.8 in 2.7, we have

Methods

(2.9)

Po,

Po,

The slope of this pressure curve is


d

PIP^,)

= 2 (1-P) sin 6 cos 6


dS
according to Newtonian theory.
For an isentropic expansion downstream of the stagnation
point, on the blunt nose, we have

P
Po,

2+(y-1)

kf2

The Prandtl-Meyer angle for expansion from sonic flow to


supersonic Mach number M is
v

=JzJw)d $ Z ]
tan-

tan

(2.14)

The rate of change of pressure w.r.t .the Prandtl-Meyer angle v is


d ( P / P o , ) - -7 M 2 ( P / P O 2 )

dM2-1

dv

( 2.15)

Equating the expressions for the pressure gradients (Eqns.


2.12 and 2.15), noting that dv = - d6

ykf: ( P 9 /Po,)
=

2(1-P) sin 6cos 6

(M~z-I)~

where, subscript q refers to conditions at the point on the body

11

12

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

where the pressure distribution slopes are equal. The location of


this point can be easily determined once the value of Mqis known.
Eliminating 6 in Eqn. 2.16 by using Eqn. 2.1 1, we have

where

For a given free stream conditions, the value of P is known


and it is necessary to solve the Eqn. 2.17 by an iterative process.
This is done a s follows:
Assume a starting value for the matching Mach number Mq,
say 1.30. For this value calculate Q using Eqn.2.18. Calculate P
from Eqn. 2.17. Assume a new matching Mach number say 1.70
and repeat the above steps to get a new value of P. A linear
interpolation between these two calculated values is made to get
a new matching Mach number corresponding to the actual value
of P given by the free stream conditions (Eqn. 2.6). This process is
repeated until the solution converges. The location of the
matching point is easily determined once Mq is known.
From Eqn. 2.11

The pressure at the matching point, in terms of the free


stream static pressure and the ratio P is simply

Starting from the matching point, the pressure on the body


surface downstream is calculated by the Prandtl-Meyer
theoretical relationship. It is found that the use of Prandtl-Meyer
relations from the matching point onwards gives a betfer
correlation with experimental data and exact theories than by
using the sonic point a s the starting point for use of the PrandtlMeyer relation.

Methods

2.6

TANGENT WEDGE 8s TANGENT CONES

Although not based on any theoretical grounds, it is found


that the tangent wedge and tangent cone methods give
reasonably accurate results at hypersonic speeds. Tangent
wedge theory determines the pressure at each point by
calculating the pressure on a wedge of the same half angle a s the
local inclination angle at the point. The pressure on the
equivalent wedge is found by using the oblique shock theoretical
relations at the free stream Mach number of Ma.
In a similar manner, tangent cone method uses an
equivalent cone at each point to calculate the pressure on
axisymmetric bodies.
It is found that the tangent wedge method works well for
airfoils with sharp leading edges and the tangent cone method for
bodies with pointed noses.

TANGENT WEDGE

Figure 2.2. Tangent wedge method

Figure 2.3. Tangent cone method

13

14

Aerodynamic Predictive Methods and their Validationin Hypersonic Flows

When applying the tangent cone method to bodies of


revolution or equivalent bodies of revolution, sometimes it is
convenient to use an empirical equation for the pressure
coefficient rather than the use of Sims conical flow tables. The
suggested equation7 for the pressure coefficient is

which is a function of the Newtonian impact angle 8 and a so


called effective Mach number normal to the shock, Mns.The angle 6
is defined a s the smallest angle between the free stream direction
and the tangent to the vehicle surface at the point of interest. The
above equation is a physical representation of Cp for
%-dimensional oblique theory when the actual Mach number
normal to the shock is employed with Y = 1.4. The suggested
effective Mach number Mnsis

M,,

(0.87~
-~
0.554) sin6

0.53

(2.22)

which is only a function of free-stream Mach number and


Newtonian impact angle. For impact angles up to 30, the
deviation from the Sims tabulated conical flow table values is less
5% when the above expression is used for all Mach
than
numbers above 1.5.

TANGENT WEDGE, TANGENT CONE 8a DELTA


WING EMPIRICAL METHOD
Experiments on large surfaces of blunt, highly swept delta
wings in hypersonic flows have shown the following trend. For
angles of attack between 5" and 15" the tangent wedge theory
appears representative of the mean data. For angles of attack
above 15", the flow appears to change in nature such that the
tangent cone approximation appears valid up to 40" angle of
attack. Probably based on this, an approximate method has been
reported by Gentry, et. aL8,the details of which are a s follows:
For a wedge, the shock angle is given by
2.7

sin 0, =

(1 - E )

sin 8,
COS (0, - 6, )

Methods

and for a cone (thin shock layer approximation)


sine,

where,

E =

(P

sin 6 ,
= ,

,/ P ), the density ratio across the shock given by

In the limit as M ,

-+

co

The limiting values of the shock angles are:


y+1
2

-sinti,

For a wedge:

sine,

For a cone:

sin 0 ,= ---- sin 6 ,

2 6 +I)

Y +3

From Eqns. 2.23 and 2.24


Far a wedge

and, for a cone

The parameter (8 -6 ) is approximately constant and


independent of the Mach number indicating that Mm is a function of

M sin F only. For calculation purposes we need a relationship


between Mnsand M rsin F that satisfies the following requirements:

15

16

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

(a)

Shock detachment is neglected

(b)

A t M s i n g = O , MrlS=l

(c)

The solution asymptotically approaches the M,

(d)

Has the correct slope, d ( ~ ,

4%

+ cc

value

a t MI; sin& = 0

These conditions lead to a relationship of the form

M,,

K,,M,sind,
for a wedge, and
=

M,, = K , M , sins,
for a cone, where

K,

y+l

and
3

e -(K,+w,,

e sin&)

KC=

sins,)

/2

2 (Y + 1)

(Y + 3)

The pressure coefficient may now be obtained by the


following relationships for a wedge and a cone respectively,

As mentioned earlier, a t low angles of attack, the centre


line pressure distribution on a delta wing agrees well with
the 2-Dimensional theory (wedge flow) and a t higher angles of
attack with the conical flow theory. A s such, the relationships for
Mns a s given above for a wedge and a cone can be combined to
yield a relation

The static pressure jump across an oblique shock is given by

Methods

It has been observed that the value of Cp a s given by the


above expression with the value of Mns given by Eqn. 2.33
correlates well with the experimental data of pressure
distribution on the centre line of delta wings.

OSU BLUNT BODY METHOD


The Ohio State University (OSU) Blunt Body Empirical
equationg predicts the pressure distribution around circular
cylinders in supersonic flows. The suggested expression is

2.8

where p, is the surface pressure, P o , is the stagnation pressure


through the normal shock and 0 is the peripheral angle on a
cylinder (= 0 at the stagnation point). The pressure coefficient is
given by

where p is the free stream static pressure and ( P 0 , / P ,) from


the normal shock relations.

2.9

HANKEY FLAT SURFACE EMPIRICAL METHOD


This method is mainly used for estimating the lower
surface pressure on blunt flat plateslO.It approximates tangent
wedge a t low impact angles and approaches Newtonian method
at high impact angles.
The pressure coefficient is given by

DAHLEM-BUCK EMPIRICAL METHOD


The suggested Dahlem-Buck
E;mpirical metgod"
approximates the tangent cone at low angles of attack and
Newtonian at high angles.
For impact angles up to 22.5", the pressure coefficient is given by

2.10

17

18

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

C, =

[I

+ ~in(46)~/~]sin(6)~/~
(2.39)

[4 cos 6 cos (26)] 3'

Above 22.5", the pressure coefficient is given by Cp = 2 sin2 6

2.11

BLAST WAVE PRESSURE INCREMENTS

Bluntness of the body and the lifting surfaces gives rise to


a n over pressure. This additional amount must be added on to the
pressure calculated by the various methods like tangent wedge,
tangent cone, Newtonian, etc. According to the blast wave solution
given by Lukasiewicz and quoted by Gentry, et. aL8, the pressure
distribution downstream of the nose a s a function of x is given by

where
C,
d
x

is
is
is
and the values
a s follows:
Flow

2.12

the nose drag coefficient


the nose diameter or thickness and
the distance from the nose stagnation point
of coefficients A, B and nose drag coefficients are
CI

SHOCK EXPANSION THEORY

2.12.1 First Order Theory


For bodies flying a t high supersonic speeds, Eggers,

et all2-l4 and, Savin15 proposed a theory known as the


Generalized Shock Expansion theory. I n this, the flow
parameters immediately downstream of sharp nosed bodies are
calculated using either the oblique shock relations for 2dimensional bodies or the conical shock relations for ,axisymmetric bodies. Downstream of the leading edge, the body
surface is replaced by a tangent body composed of conical
segments a s shown in the Fig. 2.4.

Methods
,LEADING
TANGENT BODY
COMPOSED OF
CONICAL
SEGMENTS

EDGE SHOCK

SHOCK WAVE

Figure 2.4. First order theory


The change in the local surface slope in going from one
tangent segment of the body or the airfoil to another tangent
segment is determined and for this change, the Prandtl-Meyer
relations are used to calculate the flow properties. It is assumed
that the pressure is constant along each segment. Inherent in this
theory is that the expansion waves created at each change of slope
are absorbed by the shock and are not reflected back. Since the
theory assumes that the pressure is constant along each conical
tangent elements of the surface, the body should be slender or else
one has to consider a large number of elements to obtain accurate
pressure prediction.
2.12.2 Second Order Shock Expansion Theory (SOSET)
The previously outlined first order theory was extended
by Syvertson, et all6 by defining the pressure along a conical
frustum by a relation

instead of constant pressure along each segment. In the above


expression, pc is the pressure on the conical segment, a s given by
the conical flow over a cone of half angle equal to the slope of the
conical segment with respect to the body axis of symmetry. p, is
the pressure just aft of the conical segment a s calculated from the
Prandtl-Meyer relation from known values in region 1 (Fig. 2,5).

The pressure gradient is approximated by the relation16

19

20

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows


MACH LINES

Figure 2.5. Flow about a frustum element

where, r is the radial coordinate in the (x,r,p ) cylindrical


coordinate system with origin at the nose, and

For negative angles such a s would occur on a boat tailed


body, pc is replaced by p , . If q becomes negative, the Second
Order theory is replaced by the First Order theory. This is
because the Eqn 2.41 will not give the correct asymptotic cone
solution for negative values of q .
It has been observed that the Second Order Shock
Expansion theory predicts fairly accurately the pressure
distribution on the surface of the body at low to moderate angles of
attack and the Mach number greater than 2.0.

22

Aerodynamic Predictive Methods and their Validation in HypersonicFlows

The above relations give the coordinates of the


streamlines in the wind axis system and these coordinates are
used to generate bodies of revolution for each radial angle 4 of
interest.
Y

Figure 2.6. The axis system

Bodies of revolution thus generated for several angles of


are shown below.

EQUIVALENT BODIES

Figure 2.7. Typical equivalent body shapes

The zero angle of attack method (Newtonian plus second


order shock expansion) is applied to the equivalent bodies thus
generated from the transformed streamlines to get the pressure
distribution. There are two limitations in this method. First, the
angle of attack should be low enough so that the stagnation point
remains on the spherical surface and secondly, the angle of attack

24

Aerodynamic Predictive Methods and their Valzdationin Hypersonic Flows

AC,

=-

( 2 a )s i n ( 2 ~ ) s i n ( ~ ) + ( ~6c) oa's ~

[ ( 4 / 3 ) sin(26) sin

(o)]
. . a'

(2.49)

where

AC P = -

(2a) sin (26) sin(@)


3

(2.50)

Eqn 2.49 is used for pointed body configurations as well a s for


blunt body configurations in the windward plane area, 60" c 4 5 180".
For the leeward plane area on blunt bodies Eqn 2.49 is replaced by
Eqn 2.50. In the above equations p = (M2- 1)l! ; is the local surface
slope of the body with respect to body axis and 4 is the position on
the body surface with 4 = + 90' being the vertical plane and 4 = - 90"
corresponds to leeward plane.
2.14 VAN DYKE UNIFIED THEORY
A method based on the hypersonic small disturbance
theory applicable to both the supersonic and hypersonic flow
regimes was proposed by Van Dyke and is known a s the Unified
Supersonic Hypersonic Flow theorylg.The det.ails of the derivation
based on the similarity conditions are given by Shapiro20. The
next section also gives the derivation of the pertinent equations
which are further simplified. For small deflection angles a t high
Mach numbers, the pressure coefficient on a compression
surface is given by

where, H i s the hypersonic similarity parameter given by H = ' M6,


and 6 is the thickness ratio. The above relation can also be
applied to supersonic flows if the hypersonic similarity parameter

Mci is replaced by the factor

f i 6 a s suggested by Van Dyke.

Methods

25

A similar analysis has been applied to the high Mach


number flow on a surface in expansion flow with no leading edge
shock wave a s in the case of a flat plate at an angle of attack. The
resulting expression20 is

A s before, the similarity parameter H =


applicability in both supersonic and hypersonic flow
2.15

m1
S for

2-D AIRFOIL THEORY IN HYPERSONIC FLOWS

It is possible to derive a much simpler expression for the


surface pressure coefficient than those given by Eqns 2.5 1 and 2.52
above, based on additional assumptions, justifiable in hypersonic
flows over thin 2-dimensional bodies at small angles of a t t a ~ k ~ l - ~ ~ .
Details of the analysis are
the pressure coefficient C

,- Y M :

P-Pm
m

Let the subscript 2 denote the conditions immediately


behind a n oblique shock wave, there follows

From the oblique shock relations we have


2

Y + l

sin P = 4 CP2

and

(c~2;"")2

[(

1 - -'i2

where p is the shock angle and

)
e

sin P tan

e2

I'

the flow deflection angle.

For thin airfoils in hypersonic flow, we can approximate a s


follows:

Methods

The Prandtl-Meyer expansion is isentropic, hence we have


the relation,

Combining the Eqns. 2.58 and 2.59 2nd neglecting higher


order terms we have

The above is equivalent to

The surface pressure coefficient is

Making use of the Eqns. 2.53 and 2.61, the above can be
expressed as

Substituting for
becomes

'pW2

from Eqn. 2.57, the above equation

27

28

Aerodynamic Predictive Methods and their lralidation in Hypersonic Flows

The above equation leads to two cases viz., no shock and


no expansion.
Let us consider the no shock case. For this Ow,

0.

Hence, Mw?= bfw

The above equation can be used to calculate the


pressure coefficient on the upper surface of a flat plate at an
angle of attack of rr=8,. The case of no expansion after the
shock implies that Q

0.Hence,

The above corresponds to the lower surface of a flat plate


at an angle of attack

ow, = a .

The lift coefficient of an airfoil of chord c is given by

For the flat plate case, at an angle of attack of a , the lift


coefficient can be obtained by the use of Eqns. 2.65 and 2.66, viz.,

30

Aerodynamic Predictive Methods and their Validation i n Hypersonic Flows

A binomial expansion of the square root term of Eqn 2.69


results in

(2.72)
The above Eqns. 2.71 and 2.72 are identical up to the first
two terms and differs from each other in the third term by 10 per
cent. It is reasonable therefore to assume that the Eqn. 2.7 1 is
applicable to both the shock and expansion processes. This
assumption is equivalent to neglecting entropy change across the
shock. Both compression and expansion are considered a s
isentropic. The pressure coefficient at any point on the surface of
an airfoil is given by

Where, E is the thickness ratio of the airfoil and K E M, E the well


known hypersonic similarity parameter.
The above equation can be integrated for most of the thin
airfoil shapes to give closed form solutions for the aerodynamic
coefficients for various airfoils. Results of these calculations for
the determination of aerodynamic characteristics of various types
of airfoil shapes commonly encountered are g i ~ e n ~and
' . ~they
~ are
reproduced in section 3.2.5
2.16

HIGH MACH NUMBER BASE PRESSURE

The base region of a hypersonic vehicle will invariably be


in the shadow region. Further at hypersonic Mach numbers the
expansion of the flow from the body surface to the base region will
be such that the base portion will be in a vacuum environment. For
this condition the pressure coefficient at the base is given by:

However, due to viscosity and real gas effects some


pressure is felt in the base region and according to some
experiments for air this value is approximately 70 per cent

Methods

vacuum. Based on this, the base pressure coefficient can be


taken as

REFERENCES
Weibust, Erling. Status report on the FFA version of the
missile aerodynamics program LAIZV, for calculation of
static
aerodynamic
properties
and
longitudinal
aerodynamic damping derivatives FFA. The Aeronautical
Research Institute of Sweden, Stockholm, 1981. TN-AU1661.
Hankey, Jr., W.L. & Alexander, G.L. Prediction of
hypersonic aerodynamic characteristics for lifting vehicles.
WPAFB, Ohio, September 1963. ASD-TDR-63-668.
Sieff, A. Secondary floml fields embedded in hypersonic
shock layers. NASA, May 1962. TN-D-1304.
Sieff, A., & Whitting, W.E. Calculation of flow fields from
bow-wave profiles for the downstream region of blunt-nosed
circular cylinders in axial hypersonic flight. NASA, 1961.
TN-D- 1147.
Maslen, S.H., & Moeckel, W.E. Inviscid hypersonic flow past
blunt bodies. J. of Aero. Sci.,1957, 24(9), 683-89.
Kaufman-11, L.G. Pressure estimation techniques for
hypersonic flows over blunt bodies. J. of Aero. Sci., 1963,
lO(2).
Pittman, J.L. Application of supersonic linear theory and
hypersonic impact methods to three nonslender hypersonic
airplane concepts at mach numbers from 1.10 to 2.86.
NASA, December 1979. TP- 1539.
Gentry, A.E.; Smyth, D.N. & Oliver, W.R. The Mark IV
supersonic-hypersonic arbitrary-body program. WPAFB,
Ohio, November 1973. 11 p. AFFDL-TR-73-159.
Gregorek, G.M., Nark, T.C. & Lee, J.D. An experimental
investigation of the surface pressure and the laminar
boundary layer on a blunt flat plate in hypersonic flow,
Vol 1. March 1963. ASD-TDR-62-792.
Hankey , J r ., W.L. Optimization of lifting re-entry Gehicles.
March 1963. ASD-TDR-62- 1102.

31

32

Aerodynamic Predictive Methods and their Validatiott in Hypersonic Flows

Dahlem, V. & Buck, M.L. Experimental and analytical


investigations vehicle designs for high lift-drag ratios in
hypersonic flight. June 1967. AFFDL-TR-67- 138.
Eggers, A.J.; Sjvertson, C.A.; & Kraus, S.A. A study of
inviscid flow about airfoils a t high supersonic speeds. NACA
Report, 1953. TN-1123.
Eggers, A.J. & Savin, R.C. A unified tw-o-dimensional
approach to the calculation of three-dimensional
hypersonic flows with applications to bodies of revolution.
NACA Report, 1955. TN- 1249.
Eggers, A.J. & Savin, R.C. Approximate methods for
calculating the flow about nonlifting bodies of revolution a t
high supersonic airspeeds. NACA, 1951. TN-2579.
Savin, R.C. Application of the generalized shock expansion
method to the inclined bodies of revolution travelling at
high supersonic airspeeds. NACA, 1955. TN-3349.
Sqvertson, C.A. & Dennis, D.H. Second order shock
expansion method applicable to bodies of revolution near
zero lift. NACA, 1957. TR-1323.
Jackson, C.M.; Sawyer, W.C.& Smith, R.S. A method for
determining surface pressures on blunt bodies of
revolution at small angles of attack in supersonic flows.
NASA, 1968. TN P-4865.
Dejarnette, F.R.; Ford, C.P. & Young, D.E. A new method for
calculating surface presures on bodies at an angle of attack
in supersonic flow. Paper presented a t AIAA 12THFluid and
Plasma Dynamics Conference, July 1974, Williamsburgh,
Va., AIAA Paper No. 79- 1522.
Van Dyke, M.D.A study of hypersonic small-disturbance
theory. NACA, 1954. Report No. 1194.
Shapiro, A.H. The dynamics and thermodynamics of
compressible fluid flow. The Ronald Press, Vol. 2, 1953,
pp. 753-754.
Kaufman-11, L.G. & Scheuing, R.A. An introduction to
hypersonics. Grumman Aircraft Engineering Corporation,
August 1960. Research Report RE-82.
Linnel, R.D. Two-dimensional airfoils in hypersonic flows
JAS, 1949, 16(1).
Dorrance, W.H. Two-dimensional airfoils at moderate
hypersonic velocities. JAS, 1952, 19(9).

CHAPTER 3
AERODYNAMIC CHARACTERISTICS OF
VEHICLE COMPONENTS
3.1

INTRODUCTION

Having identified the method or methods to calculate the


local pressure on an elemental area of a vehicle component based
on its geometry, the forces and moments experienced by that
component can be obtained by integrating the pressure and
moment over the entire surface. To determine the aerodynamic
characteristics of the complete vehicle, it is the usual practice to
divide the general configuration into simple basic components
such as nose, body, lifting surface, control surface, etc. By
summing the aerodynamic characteristics of the isolated
individual components and the effect due to the interference of
one component on the other, the complete vehicle characteristics
are determined. For example, the normal force coefficient of a
vehicle can be expressed as
C N = C N B + C N W + C N T + C N B (W ) + C N B (T ) + C N W ( B ) + C N T ( B ) +
C N T (W ) + C N W (T )

The subscripts B, W, and T refer to the body, exposed wing


and exposed tail, respectively. The terms CNB (W ) , CNW (B) , etc.
represent the interference effects of the exposed wing on the body,
of the body on the exposed wing, etc.
The forces and moments are normally expressed in
coefficient forms referred to either in the body-fixed coordinate
system (the normal CN and axial CA, force coefficients) or the wind
oriented coordinate system (the lift (CL ) and drag (CD ) force
coefficients). These can be converted from one system to the other
by the following relations.

34

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows


Z

CN
CL

CD
cm
y

x
CA

Figure 3.1. Connection between wind-oriented coefficients (CL, CD)


and body-oriented coefficients (CN, CA).

CL = CN cos CA sin
CD = CA cos + CN sin
CN = CL cos + CD sin
CA = CD cos CL sin
3.2

BODY-ALONE AERODYNAMICS

3.2.1 Forces & Moments on the Body


The bodies of hypersonic vehicles usually have a blunt nose
followed by a conical frustum and a cylindrical after body. Some of
the vehicles may have an ogival nose followed by a cylindrical body.
Forces on body shapes of these types can be analysed using
Modified Newtonian theory (MNT) or the Second-order Shock
Expansion theory, coupled with cross flow drag analysis on
cylindrical portion. Some illustrative examples using MNT are
worked out.
According to the MNT, the coefficient of pressure on
surfaces exposed directly to the flow 1 is given by
C p = K (sin cos sin cos sin ) 2

(3.1)

where
= Angle made by surface of body with body axis
= Angle of attack of the body axis
= Polar angle of any point on body surface, measured
from positive xy plane and positive for counterclockwise
direction when viewed from rear.

Aerodynamic Characteristics of Vehicle Components


Z
ZN

SHIELDED AREA

tan
u = sin 1

tan

Figure 3.2 Modified Newtonian theory

In the Newtonian theory, the pressure coefficient C p u in


the shielded region of the body is zero. However, from gasdynamic considerations, the value of C p u lies between zero and

2 / M 2

depending on the free stream Mach number, shape of

the body, and angle of attack. Experiments have indicated that


the largest negative value of C p u appears to be of the order of
1/M 2 for

= 1.4. Therefore, for high Mach number flows, it is

reasonable to assume that the shielded regions do not contribute


to the forces and moments.
For any portion of the body shown in Fig. 3.2, the axial
force is given by
/2
u

A = 2q
ds
C p r sin d+ C p u r sin d

surface / 2
u

Cp as given by the Eqn. 3.1, C p u = 0 and ds = dx/cos

(3.2)

35

36

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

A = 2q K

r tandx

(sin cos sin cos sin)

A
Axial force coefficient, CA = q S
ref

2K
CA =
Sref

dC A
dx

r tan dx

(sin cos sin cos sin ) d


2

/ 2

length

S ref

(3.4)

2 Kr

(3.3)

/2

length

tan sin 2 cos 2

u
sin 2 u

+
2
4
4

(3.5)

u +
+ cos 2 sin 2
2

+ 2 sin cos cos sin cos u

(3.6)

Similarly, the normal force is given by

/2
u

N = 2q
dx
C p r sin d + C p u r sin d
2

u
length
/

(3.7)

with Cp given by the Eqn. 3.1 and C p u assumed zero,

dC N
dx

Kr
S ref
+

sin 2 cos 2 u + tan


2

cos u cot tan 2 + 2 tan


3

(3.8)

Likewise, the pitching moment, taken about the centroid


of the area of the base of the nose section, for convenience, is
given by

Aerodynamic Characteristics of Vehicle Components

M = Kq
{(L N x ) r tan }dx

length

C p r sin d +

/ 2

dC m
dx

Kr
S ref L

C p u r sin d

u

/2

(3.9)

sin 2 cos 2 [(L N x ) r tan ]

2
u + tan + cos u cot tan + 2 tan
2
3

(3.10)
where, L is the reference length. Eqns. 3.6, 3.8 and 3.10 can be
integrated analytically or numerically to give axial force, normal
force, and pitching moment coefficients, respectively for any
arbitrary shaped body of revolution.
3.2.2 Axial Force
The axial force coefficient on a body can be considered to be
made up of three parts,
C A =C A f +C A b +C A N

(3.11)

where
CAf

= Coefficient of axial force on the body due to skin friction

C Ab

= Coefficient of axial force due to the pressure on the base


area, and
= Coefficient of axial force on the body excluding friction.
(wave drag)

C AN

3.2.3 C A f Skin Friction Coefficient


Several methods have been presented in the literature for
the prediction of skin friction on bodies and flat plates in
supersonic and hypersonic flows. Majority of these are complex,
laborious and require a detailed knowledge of the flow field over the
body for the evaluation of the skin friction coefficient. However, in
preliminary design analysis, it is sufficient to go for simple

37

38

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

methods which give fairly acceptable values. Towards this, the


incompressible flow results are utilised with an empirical
correction for compressibility or Mach number effects.
Axial skin friction coefficient C A f = C f

S B wet

(3.12)
S ref
The mean value of Cf depends on whether the flow is
laminar or turbulent. It is an usual practice to assume that the
boundary
layer
over
the
body
is
laminar
if
6
Re < Re cr = 10 (Reynolds number based on the length of the
body). For laminar flow
Cf =

1.328 C f
Re C f i

lam

(3.13)

The above is the well known Blasius relation for the flow
over a flat plate in incompressible flow multiplied by the
compressibility factor. The suggested value of the compressibility
factor is
Cf

Cf

= 1 0 .028 M

lam

(3.14)

An alternate expression suggested for Cf in laminar flow


is
C

1
Re

[1.328 0.0236 M

+ 0.000349 M

0.00335 M

8.54 M

(3.15)

For Re > Re cr = 10 6 the boundary layer is turbulent with a


laminar part in front of it. For this case the value of Cf is given by

Re Recr
C f = C fi

Re




lam

(3.16)

C f i = 0.427 (ln Re 0.407 ) 2 . 64 3300 Re 1

(3.17)

Cf

Cf
i

Recr

Re
turb

Cf

Cf
i

where

C f

= 1 0.0689 M 0.0343 M 2 + 0.0061 M 3 0.000278 M 4


C
f i turb
(3.18)

Aerodynamic Characteristics of Vehicle Components

The above compressibility factors are for the case of an


adiabatic wall. The skin friction coefficients (given above) both for
laminar and the turbulent flows are for a flat plate at zero angle of
attack. As the actual vehicle has a finite thickness, the pressure
gradient and the boundary layer displacement effects influence
the skin friction. To account for this, Hoerner2 has suggested that
the skin friction coefficient values as given above be multiplied by
the factor
d

1 + 1.5
L

3
2

d
+ 7
L

(3.19)

3.2.4 C A b Base Pressure Coefficient


The base drag of a hypersonic vehicle may be a substantial
part of the total zero lift drag. Based on experimental data, various
empirical formulae have been suggested. A few of these formulae
are listed below. It is to be noted that some of the formulae are
expressed in the form of axial force coefficient and some in the
form of base pressure coefficient.
C D

base

C P

C p base =

Cp

C Ab

cyl .

S base

S ref
1
M

1. 4

2
=
2
M +1

base

base

0 . 57
M 4

2 .8

for M 1

(3.203)

2 M 2 ( 1 )

+
1
(
)

(3.214)

= (1/cos )

(0.004714 M 2 0.06307 M + 0.2455 ) 4Sd

(3.225)

ref

C Aboattail = (0.0071 M + 0.782 ) C Ab

cyl

(3.235)

It has been experimentally observed that for M 5.5 there


is very little effect of angle of attack on the base pressure. It is
well known that as the Mach number becomes very high, the base
drag coefficient approaches zero.

39

40

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

3.2.5 Determination of C AN , the Axial Pressure Coefficient


of Nose Portion of the Body
3.2.5.1 Pointed Cone
Calculations for a conical nose are simple. Since = v, the
semi-vertex angle of the cone is constant and r is a linear function
of x, (dr = dx tan). For positive angles of attack less than v, u = /
2 and all surfaces of the cone are exposed to the flow. For v
v, certain portions of the cone are shielded from the flow
and for this case u = sin 1 (tan v / tan ). For v , none
of the surfaces of the cone are exposed to the flow and u = /2.
For the above range of angles of attack, the axial pressure
coefficients of the cone are

C Acone =

Kd 2
sin 2 cos 2
2
2

cos sin +
4S ref
2

; 0 v
(3.24)

where, d is the base diameter of the cone. If the axial force


coefficient is based on the base area, then Sref = ( d2/4). The axial
force coefficient is

CA

cone

= K cos 2 sin 2 +

sin 2 cos 2
2

; 0 v

(3.25)

For the case, v , the upper limit of the integration


u = sin 1 (tan /tan ) .

CA

=
cone

2
2
cos sin u +
2

sin 2 cos 2

u +
2
2

sin 2 sin 2 cos u


+
8

and
C Acone = 0

for v

(3.26)

Aerodynamic Characteristics of Vehicle Components

3.2.5.2 Pointed Ogive

C Aogive

= K 2 1 + F 2

[1 + 0.22 F

1+ F2
1 F 2 ln

F2

sin 2 M 2 1

)] 4Sd

(3.27)

ref

where, F = Ln /D is the fineness ratio.


3.2.5.3 Hemispherical Nose
For the body having a hemispherical nose, the axial force
coefficient is given by

CA

=
hemisphere

K R2

(1 + cos )2

4 S ref

(3.28)

3.2.6 Normal Force


3.2.6.1 Pointed Cone
When the Eqn. 3.8 is integrated for the case of a right
circular cone having a semi-vertex angle of v , and base diameter
d, the following expressions result for the normal force coefficients:

C N cone = K

C N cone =

K
2

cos 2 v sin 2 d 2
2

4 S ref

; 0 v

(3.29)

cos 2 v sin 2

2
1

d
u + 2 + 3 cos u (cot tan v + 2 cot v tan ) 4 S

ref

: v v
CN

cone

=0

(3.30)

(3.31)

41

42

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

3.2.6.2 Pointed Ogive

CN

ogive

d2

1+ F2
= K (sin cos ) sin F 2 2 1 + F 2 ln

F2
4 S ref

(3.32)
It has been found that for angles of attack near zero for a
circular arc ogival nose, the normal force can be adequately
obtained from the simple equation for that of the cone of equal
fineness ratio. For fineness ratios of unity or larger, calculations
show that the difference in C N for the cone and ogive are negligible
at angles of attack up to v for the cone. At angles of attack
somewhat less than v of the cone, the curves of C N versus for
the two nose shapes cross, so that = v , C N for the ogive exceeds
that for the cone. No explicit expression for the location of the
centre of pressure can be given for the ogive as in the case of the
cone; computations have shown, however, that for small angles of
attack the centre of pressure of the ogive is nearer the vetex than
that of the cone of equal fineness ratio and moves rearward with
increasing angle of attack.
3.2.6.3 Hemispherical Nose
Integration of the Eqn. 3.8 for the case of hemispherical
nose results in

C N sphere =

R 2 K
4S ref

sin (cos + 1 )

; 0 ;

(3.33)

Calculations for other nose shapes are more involved


since r, and u are all functions of the lengthwise variable x. In
general, closed form solutions cannot be obtained and one has to
go for numerical integration.
3.2.6.4 Cylinder
The normal force contribution from the circular cylindrical
portion of the fuselage can also be obtained from Eqn. 3.8. If the
normal force coefficient is based on the base area of the nose
which is the same as the base area of the cylinder, the expression
for the normal force coefficient is

Aerodynamic Characteristics of Vehicle Components

CN

=
cylinder

1.5
d (L B L N
S ref

) sin 2

; 0 (3.34)

In the above expression d is the diameter of the cylinder


and a value of 2 is substituted for K, the factor in the modified
Newtonian expression for the pressure coefficient. The normal
force coefficient as given by the above agrees very well with the
experimental data at small to moderate angles of attack and
overestimates the force by only 5 per cent near = 90.
3.3

ALLEN & PERKINS VISCOUS CROSS FLOW


THEORY

In cases in which the force on a body such as a fuselage is


determined by some other method other than the Newtonian
method, such as Second Order Shock Expansion method, Hybrid
theory of Van Dyke, etc., then it is necessary to include the
contribution to the lift by viscous cross flow. The widely used
method to calculate the lift due to viscous cross flow is that
proposed by Allen and Perkins6. According to this theory which is
fairly simple yet quite powerful, the inviscid and viscous effects of
the flow are assumed to be separate, with the inviscid form of the
solution applying to the axial flow while the viscous part is confined
to the cross flow giving rise to a nonlinear lift force coefficient. The
viscous cross flow contribution to the lift coefficient is given by

( CN )NL

= C dc

Ap
Aref

sin 2 cos

Here, is the drag proportionality factor or cross flow drag of a


cylinder or flat plate of finite length to one of infinite length, (for
high Mach number flows, a value of =1 is normally used). C d c is
the average cross flow drag coefficient, A p is the planform area of
the body in the cross flow plane and A re f the reference area used in
lift force coefficient term. C d c is taken from the experimental
section drag coefficient. For simplicity, in the absence of available
experimental data, a value of about 1.2 is normally assumed.
3.4

MOMENTS

The pitching movement is caused by the normal forces


acting on the body nose and the body cylindrical part.
Cm = C

+ Cm
N

cylind er

43

44

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

3.4.1 Pointed Cone


From the integration of the Eqn. 3.10, where the moment is
taken about the centroid of the base of the nose and taking the
length of nose as the reference length, the pitching moment
coefficient, for the case of a cone is given by

C m cone =

Cm

cone

(1 2 tan 2 v )cos 2 v sin 2

0 v

K
2
2
2
1 2 tan v cos v sin
6
1

u + 2 + 3 cos u cot tan v + 2 cotv tan

v v

C m cone = 0 ;

(3.35)

);

(3.36)

(3.37)

It is seen from the above results that the location of the


centre of pressure of the conical nose is independent of the angle
of attack. Its distance from the vertex is given by

x cp
LN

2
1
3 cos 2 v

(3.38)

3.4.2 Pointed Ogive


CN
ogive
C m ogive =
(Lmr - l N ) + 3K (sin + cos )sin F 2 1 + F 2
L ref

1 F arctan
F

2
d lN
1 d

3 4 S ref L ref L ref

)
(3.39)

where
F
Lmr
L re f

=
=
=

( L n /d) the fineness ratio


the moment arm length measured from the nose
the reference length in the moment coefficient
expression

No explicit expression for the location of the centre of


pressure can be given for the ogive as in the case of the cone;
computations have shown, however, that for small angles of attack

Aerodynamic Characteristics of Vehicle Components

the centre of pressure of the ogive is nearer the vertex than that
of the cone of equal fineness ratio and moves rearward with
increase in angle of attack.
3.4.3 Hemisphere
When the moment is taken about the base of the sphere,
C m hemisphere = 0 ;

(3.40)

3.4.4 Circular Cylinder


The assumption is made that resultant normal force acts
through the mid-point of the cylindrical body, thus giving

Cm
3.5

=
cylinder

CN

cylinder

L ref

[L mr (l N + l cylinder )]

(3.41)

WING ALONE AERODYNAMICS

At hypersonic Mach numbers it is assumed that the


different parts of the vehicle such as body, wing and rudder have no
influence on each other. The wing or rudder is considered in
isolation. Most of the engineering methods available to analyse
the pressure over a lifting surface are based on 2-dimensional
theory. In some cases it is possible that wing tips might have an
appreciable effect on the predicted lift. In such cases
2-dimensional lift is corrected for 3-dimensional flow by the use
of supersonic linear theory.
The common types of wing sections utilized in hypersonic
vehicles are:
(a)
(b)
(c)

Hexagonal
Biconvex, and
Blunt nose leading edge

3.5.1 Hexagonal Shape Wing Section


The analysis is based on the formulations given by
Weibust5. In this case the wing geometry has wedge sections at
the leading and trailing edges with straight portion in between and
treated as having six surfaces. The oblique shock and PrandtlMeyer theories are applied to the leading and trailing edges
depending on whether the surface under consideration gives rise
to compression flow or expansion flow. For forces normal to the
chord the wing is treated as a flat plate with zero thickness.

45

46

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Compression Flow
The pressure coefficient behind an attached oblique
shock is given by the relation,

C pc =

4 M 2 sin 1

( + 1) M 2

(3.42)

where, is the shock angle obtained from oblique shock tables


or from the solution of the cubic shock equation in sin2

sin 6 + A sin 4 + B sin 2 + C = 0


where
A=

B=

2M

+1

C =

+2

sin 2

( + 1 ) 2 1
sin 2
+
+
2

4
M

cos 2
4

where, is the flow deflection angle and it is assumed that for the
given free stream Mach number its value is such that it always
gives rise to an attached shock at the leading edge.
Expansion
The pressure coefficient behind an expansion fan is

C pe =

2
M 2

pe

1 =
M 2
p

pe
p
o

po
p

(3.43)

Where Pe is the pressure behind expansion obtained either from


P-M expansion flow tables or from the following relations
poe
1 2
M
1 +
p
2

from the isentropic relation (3.44)

Aerodynamic Characteristics of Vehicle Components

and

pe
poe

1
=
1 + cos
+1

1
e + arctan
+1

2
e

(3.45)
where, pe is the pressure and Me the Mach number downstream
of the expansion. e is the corresponding Prandtl Meyer angle
which is obtained from the Prandtl-Meyer function

e =

+1
1

arctan

1
+1

(M

2
e

1 arctan M

2
e

(3.46)

similarly

+1
1

arctan

1
+1

(M

1 arctan

(3.47)
and, e = +
From the known value of e the Mach number Me is
obtained by an iterative solution of the Eqn. 3.46, then from pe
Eqn. 3.45 and C p e from Eqn. 3.43.
3.5.1.1 Axial Force
The axial force on a wing surface consists of two parts, one
due to friction and the other due to inviscid flow over it. The friction
force is determined similar to the case of a body in hypersonic flow
(section 3.22).
CAf = C

S wet
f

S ref

(3.48)

In case the vehicle has horizontal and vertical wing


surfaces (cruciform arrangement as in a missile) the axial force is
the sum of these two. The wing surfaces might also act as rudders
by wing deflection (see Fig. 3.3, 1& 2). The pressure forces on the
leading and trailing wedge surfaces are calculated by the

47

48

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

application of appropriate oblique shock or Prandtl-Meyer


theories. The inviscid flow over the straight portion does not
contribute to the axial force.
HORIZONTAL WING OR RUDDER
WHEN = 0

CN

CA
X,X'

Z'
Z

= ROLL ANGLE

Y'
Y

RUDDER
DEFLECTION

'

VERTICAL WING OR RUDDER


WHEN = 0

Cruciform wing arrangement

1
2

EXPANSION

SHOCK

Figure 3.3 Shock and expansion method applied to hexagonal crosssectional wing.

Aerodynamic Characteristics of Vehicle Components

11
INNER WING

i ,1

SECTION-AA
D

0 ,1

OUTER WING

i ,2
A

0 ,2

21

B
SECTION-BB
bi
bo

Figure 3.4. Planform showing inner and outer wing panels

The axial pressure coefficient on a horizontal wing is given by


C AwH =

t
2 S ref

{ (b i d )

( 1)

m +1

m =1 n =1

+ (btot bi

( 1)

m =1 n =1

m +1

C p im , n , M

C p om , n , M

i m , n

i m ,n

o m , n

o m ,n

(3.49)

where t is the wing thickness, m = 1 for LE and m = 2 for TE , n = 1


for upper and n = 2 for lower surfaces, = | ' | is the deflection
angle. If ' > 0 (shock) and Cp is determined from Eqn. 3.42 and
' < 0 (expansion) from Eqn. 3.43. Let the semi-wedge angles
normal to the leading and trailing edges be denoted by 1' and 2'
respectively. Subscript i refers to inner wing, and o to the outer
wing.
The angle i for the inner horizontal wing is given by

cos cos '


i m,n = ( 1)m + arcsin sin m
im

sin
+ sin m
i

1
]
sin ' sin + ( 1 )n ( 1)m + sin ' cos cos m

(3.50)

49

50

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

where, m = 1 for LE and m = 2 for TE, n =1 for upper and n = 2 for


lower wedge surfaces, = roll angle, ' = the non-rolled angle of
attack and subscript i refers to the inner wing.
'
Similarly, o m ,n for the outer wing is as given above with
o m replacing im

The axial force for the vertical wing C A wV is also obtained


from the Eqn. 3.49. However, the deflection angle ' is obtained
by substituting with (90 - ) in Eqn. 3.50.
The wing total axial force is given by
C

=C A

+C A

wH

+C A

wV

(3.51)

In case there is deflection of the rudder by an angle


then one has to consider the contribution to the axial force by this
deflection. The change in ' due to rudder deflection angle and
hence to C A R is usually neglected. This is because the axial
component of the force normal to the rudder, C A R , is based on
the total wing planform including the areas on the wedge shaped
LE and TE.
The total axial force on the rudder is given by
C A = C A + CA
R

RH

+ CA

RV

+ C AR

(3.52)

where, C A f is obtained from Eqn. 3.48 and C A R H and C A R V are


obtained from the Eqn. 3.49 with correct values of ' . C A R from
the normal force as outlined below.
3.5.1.2 Normal Force
An assumption is made that the normal force acts through
the area centroid of the wing planform. As mentioned earlier, for
normal force calculation, the wing is considered as a flat plate
with zero thickness.
The actual wing or rudder angle of attack ' expressed in
non-rolled angle of attack ' , roll angle and rudder deflection
angle is

Aerodynamic Characteristics of Vehicle Components

j ,k = arc sin sin j ,k cos

{(

+ cos ( j 1) 90 o ( 1 ) j

) }cos

j ,k

sin

(3.53)

where, j = 1 for the horizontal wing and j = 2 for the vertical wing.
k =1 for the right and upper wing halves and k = 2 for left and
lower wing halves at roll angle = 0. When the above angle is
known the pressure coefficients are determined by use of Eqns.
3.42 and 3.43. The above gives the forces acting normal to the
wing or rudder chord. This has to be converted to get the axial
component of the force and the component of the force normal to
non-rolled xy plane ( i.e., negative z axis direction).
3.5.1.3 Axial Component of the Rudder

C AR

SR

4 S ref

j =1 k =1

C ( , M ) C ( , M )
p
j ,k
j ,k
pe

c
sin j , k

(3.54)

In the above equation SR is the planform area of either


horizontal wings (right and left combined) or vertical wings (top and
bottom combined).
3.5.1.4 Normal Component (Wing or Rudder)

CNw =

Sw
4S

ref

j =1

2
C p ( j ,k , M ) C p ( j ,k , M )

c
e
k =1

' j ,k
cos ( j 1) 90 o ( 1) j cos j , k

j ,k

(3.55)

'
with j ,k = | j ,k |, j = 1 for horizontal wing and j = 2 for vertical
wing, k = 1 for right and upper wing halves and k = 2 for left and
lower wing halves at zero roll angle viewed from nose towards tail.

It is suggested that as a means for conversion from


2-dimensional to 3-dimensional lift at hypersonic speeds, the
following approximation for wing tip effects, based on supersonic
linear theory, may be applied.

51

52

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

C L

CN =

Cn

4
M

(3.56)

where, C N is the normal force coefficient, C L is the lift-curve


slope from linear theory for the 3-dimensional planform, as given
by Harmon & Jeffreys7, and Cn is the 2-dimensional normal force
coefficient calculated using 2-dimensional shock and expansion
theory.
3.5.1.5 Pitching Moment
The pitching moment is given by

Cm

=
W

CN

X ref

(X

mr

X cp

(3.57)

3.5.2 Other Wing Sections


For wings having a blunt leading edge, the commonly used
method to analyze the pressure on the wing surface is by the
Newtonian + Prandtl-Meyer method as described in section 2.4.
For wings having a sharp leading edge giving rise to an
attached oblique shock and having curved surfaces like the
biconvex airfoil, etc., the tangent wedge method is applicable. For
thin sharp edged wing sections at low angles of attack, the
approximate 2-dimensional airfoil method described in section
2.14 can be applied to determine the pressure distribution over the
airfoil surfaces. From this, the axial force coefficient, the normal
force coefficient and the moment coefficient can be determined.
Based on methods described by Kaufman et. al 8 and Dorrance9,
the results of calculations done on some common airfoil sections
are given. The same is reproduced in the accompanying table.

Aerodynamic Characteristics of Vehicle Components

3.5.2.1 Airfoil Characteristics by 2-dimensional Hypersonic


Airfoil Theory
All equations are based on a dimensionless coordinate
system( , ) with origin at midchord.

2x
;
c

2z
;
c

c = chord length, ;

= angle of attack, K = M ;

= thickness ratio

Moment Cm taken about the origin;


A1

2
,
K

A2 =

+ 1
,
2

t = thickness;

A3 =

t
; = ;

Validity 3 M 12

+ 1) K
6

53

54

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Profile

Airfoil Characteristics

CN
= 2 ( A1 + A3 )
2
Cm
=0
2

Flat Plate

CD
= 2 ( A1 + A3 )
3

CN
= 2 A1 + A3( 3 + 2 ) ;
2

Symmetrical Double Wedge

Cm
=A2
2

CD
= 2 A1 2 + 1 + 2 A 3 4 + 6 2 + 1
3

Double Wedge

CN
= 2A1 + 2A3 2 + 3 2
2

1 1

Cm
1 + 1 1 1
= A2 3 A3

4
2
1 1 1 + 1

CD
1
1
= 2A1 2 + 1 2 + A2

2
2
3

(1 1)
1 1

(1 + 1)

1
1
+ A3
+
+ 2 4 + 6 2 1
3
3
1+
+

(
)
(
)

1 1
1 1

1
1 1

Aerodynamic Characteristics of Vehicle Components

Profile

Airfoil Characteristics

CN
3
1
= 2 A 1 + 2 A 3 2 +
2

(
2 1

Modified Double
Wedge

1)

CD
1
1
= A1 2 2 +
+

3
1 1 1 2

1
1
+ A2

2
2
(1 2 )
(1 1 )

1
1

+ 6 2
+
1 1 1 2

CN
= 2 A1 4 A2 + 2 A3 3 + 2
2

Cm
1
3

= A 1 + A 2 A3 2 + 2
2
2
2

CD
= 2 A 1 2 + 2 12 A 2
3
+ 2 A 3 4 + 12 2 + 8

CN
3

= 2 (A 1 + A 2 ) + A 3 + 2 2
2

Blunt T.E. Double


Wedge

3
1

(
2 1

2 )

1 + 1 1 + 2
Cm 1
3

= A2 ( 1 + 2 + 2 ) A3
+
2
2
4

1 1 1 2

1
1
+ A3 2 4 +
+
3
(1 1 ) (1 2 )3

Single Wedge

55

CM
=0
2
CD
1

= A 1 + 2 2 + A 2 + 3 2
3
2

+ A 3 + 3 2 + 2 4
8

56

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Profile

Airfoil Characteristics

t1

CN
8
= 2 A 1 + A 2 ( 1 + 2 )
2
3

+ 2 A 3 3 + 4 A 3 12 + 22
t2

1 = 2t 1 / t
2 = 2t 2 / t
t = t1 + t 2

Cm 4
1
= A 2 A 1 ( 1 + 2 )
3
3
2
4
A 3 13 + 32 A 3 2 ( 1 + 2 )
5

Double Circular Arc

CD
2

= 2 A 1 2 + 12 + 22
3
3

+ 8 A 2 ( 1 + 2 )

14 + 24
+ 2 A 3 4 + 4 2 12 + 22 +
5

Modified Circular Arc

(
(

)
)

CN
1 13

= 2 A1 + 2 A3 2 + 4
2
2

1 12

Cm
2

CD
3

4
A2
3

(1 )

(1 )
3
1
2
1

= 2 A1 +
2

4
3

+ 2A 3

(1 )
(1 )
(1 )
+8
(1 )
3
1

2 2
1

3
1

2 2
1

16
5

(1 )
(1 )
3
1

2 4
1

Aerodynamic Characteristics of Vehicle Components


Profile

Airfoil Characteristics

CN
2
Cm
2

CD
Symmetrical Biconvex

= 2 A1 + 2 A

2 + 4 + 2

2
2

= 2 A2 +
2

3 15

8
1

= 2 A1 + 2 +
2

15
3

16
16
+ 2 A2 8 2 +
+ 4 +
2

5
5

CN

Cm

Double Parabolic Arc

CD

CN

= 2 A1 + A 3 4 + 2
4

Cm

Blunt T.E. Double


Parabolic Arc

CD

12

2 +

)]

A2

3
1

= 2 A1 + 2 + 2 A 3

= 2 A 1 + 8 1
+ 2A 3

57

(1 + 1 ) 2

4A 2

(1 + 1 )

+ 8 1 A 3

4
2
+ (1 + 1
3

2A 1

A2

4 1 + 3 12
2

(1 + 1 ) 4

4
1
A2
3
(1 + 1

(1 + 1 )

16

4 + 8 2 +

)4

(1 + 1 ) 4

+ 4 12

[ 3 1 2 ( 1 + 1 ) 4

+ 4 1 ( 1 + 12 )] + 2

A3

(1 + 1 )

[ 4 (1 + 1

)8

Contd ...

58

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

4
+ 8 2 1 + 3 12 (1 + 1 ) + 16 14 + 2 12 + ]

CN
1

Cm
Single Parabolic Arc

CN
1

= 2A 1

2
CD

A1 +

16
A2 + 2A3 8 + 2
3

32

A2 A3 2 2 +

3
5

= 2 A1 + 2 16 A 2

128

+ 2 A 3 4 + 16 2 +

2 1

= 2 A1

2
(1 + 1 )

+
Blunt T.E. Single
Parabolic Arc

8 A2

(1 + 1 )2

2 + 12

1 3
2

(1 + 1 )

6 1 2
3

+ 2 A3

(1 + 1 )2

Cm
2

8 1 + 3 12

(1 + 1 )4
2

A1

2
32 1 (1 + 1 )

8
(1 + 1 )

A2

3 (1 + 1 )
2 A3
2
4 1 (1 + 1 )
(1 + 1 )6
3

(1 + 1 )2

1
16 + 12

5

8 (1 + )2 + 2 (1 + )4
1
1
1

Aerodynamic Characteristics of Vehicle Components

CD

1
+ 12
8

= 2 A1 2 +

(1 + 1 ) 4

8 A2

(1 + 1 )

41

(1 + 1 ) 2

3
2 1 + 3 12
8 1 1 + 12
+

1 2
(1 + 1 )2
(1 + 1 )4
2

128 1 + 212 + 14
4
5

+ 2A3 +
(
+ 1)6
1

128 1 1 + 12

(1 + 1 )6

) + 16 (1 + 3 ) +
2

(1 + 1 )4

2
1

8 1 3

(1 + 1 )2

REFERENCES
1.

2.
3.

4.
5.

6.

7.

Grinminger, G.; Williams, E.P. & Young, G.B.W. Lift on


inclined bodies of revolution in hypersonic flow. JAS. 1950,
17(11).
Hoerner, S.F. Fluid dynamic drag. published by the author.
Bonner, E.; Clever, W. & Dunn, K. Aerodynamic preliminary
analysis II, Part-I. Theory, NASA, April 1991. Contractor
Report, 182076.
Gabeaud, A. Base pressure at supersonic velocities. J. of
Aero. Sci. 1950, 17(8), 525-26.
Weibust, E. Status report on the FFA version of the missile
aerodynamics program LARV for calculation
of static
aerodynamic properties and longitudinal aerodynamic
damping derivatives FFA. The Aeronautical Research
Institute of Sweden, Stockholm, 1981. Technical Note
AU-1661.
Allen, H.J. & Perkins, E.W. Characteristics of flow over
inclined bodies of revolution. NACA, March 1951.
RM A50L07.
Harmon, S.M., & Jeffreys, I. Theoretical lift and damping in roll
of thin wings with sweep and taper. NACA, 1950, TN-2114.

59

60

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

8.

9.

Kaufman-II, L.G. & Scheuing, R.A. An introduction to


hypersonics. Grumman Aircraft Engineering Corporation
August 1960. Research Report RE-82.
Dorrance, W.H. Two-dimensional airfoils at moderate
hypersonic velocities. JAS, 1952, 19(9).

CHAPTER 4
SKIN FRICTION FORCE CALCULATION
4.1

INTRODUCTION

In the previous chapter (section 3.2.2), expressions are given


for the skin friction forces which are simple modifications of
incompressible flat plate values. Being empirical relations they are
only approximate. It may be desirable to have a much better
evaluation of the skin friction forces experienced by a body in high
speed flow. Towards this, an engineering approach is adopted and no
attempt is made to calculate the detailed skin friction distribution on
the exact shape of the body. Instead, the vehicle surface is divided
into a large number of flat surface panels in a manner that
adequately approximates the true shape. Leading edge surfaces and
the curvature are omitted. In the skin friction analysis, the number
of panels chosen to represent the body is usually much less than the
panels required in the pressure calculation analysis. For each surface
element, its normal end coordinates of the area centroid are
determined. The shear force on each surface element is assumed to
act through its centroid on the surface in a direction parallel to the
plane containing the surface normal and the free stream velocity
vector. Approximate laminar or turbulent skin friction relations are
used to calculate the skin friction force on the element. The net shear
force on the body is obtained by summing up over the vehicle. In this
type of approach the problem of determining the viscous force on the
two- or three-dimensional body is reduced to one of solving the skin
friction force on a number of constant property flows over flat plate
panels. For the skin friction calculation, the local flow properties
such as pressure, temperature, velocity and density over each of the
element under consideration are required and these are determined
from the approximate inviscid aerodynamic analysis programs such
as tangent wedge, tangent cone, modified Newtonian, Shock
Expansion, Newtonian plus P.M. expansion, etc.

62

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Compressible turbulent flow over a flat plate at an arbitrary


Mach number and plate temperature has been the subject of
investigation by many workers over a number of years. Numerous
theoretical and finite-difference computational methods have been
suggested to calculate the skin friction coefficient. Amongst these,
the commonly used engineering methods are that due to Sommer
and Short1, Van Driest2 and Spalding and Chi3. All three methods
are suitable modification of the incompressible turbulent skin
friction coefficient based on Reynolds number suitably transformed
to take into account the compressibility and temperature effects.
Brief summaries of these methods are given below.
4.2

SOMMER & SHORT METHOD

This method is based on finding a temperature T * at which


the density and viscosity for compressible flow have to be evaluated
if incompressible flow relations for zero heat transfer are to apply
for any Mach number, Me , at any wall temperature ratio (Tw/Te ), for
a given Reynolds number, Rexe. The subscript e refers to conditions
at the outer edge of the boundary layer over the surface element
under consideration and x, the distance of the centroid of the
element under consideration from the nose or the leading edge. The
calculation procedure is as follows:
Evaluate T

from the equation

T*
= 1 + 0.035 M
Te

2
e

T
w
+ 0.45
Te

(4.1)

*
from the relation
Evaluate Reynolds number R e xe

Re *xe
Re xe

*
=

e
*

(4.2)

Since the pressure is constant across the boundary layer,


*
e

Te
T

Aerodynamic Characteristics of Vehicle Components

Therefore

e
*

T
e
=
T *

Re *xe
Re xe
e

The ratio *

e
*

(4.3)

is determined from the Sutherland relation, viz.,

Te
*
T

1.5

T + 115

T e + 115

( temperatur e in o K )

(4.4)

Evaluate from the Karman-Schoenherr equation given by

0.242
C F*

= log 10 C F* Re *xe

(4.5)

or alternately from the Prandtl-Schlichting equation, viz.,

C F* =

0.46

*
log 10 Re xe

2.6

(4.6)

Evaluate CF the compressible turbulent skin friction coefficient from


the relationship

CF
C F*

4.3

*
e

Te
T

(4.7)

VAN DRIEST-II METHOD

The following is the formula for calculation of local skin


friction coefficient due to turbulent compressible flow over a flat
plate due to Van Driest2.

63

64

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

4.15 log 10 Re x C F
w

sin 1 A + sin 1 B
T
aw
CF
Te

+ 1. 7

(4.8)
where
A =

2a

2
b + 4a

0.5

and

B=

b
2
b + 4 a 2

0.5

The parameters a and b are given by the following relations



1

a =
M
2

2
e

T e
aw

1
and b =
Tw
Tw

The adiabatic wall temperature is given by the relation

Taw
Te

=1 +

1
2

M e2

where, = the recovery factor = (Prandtl No.)1 / 3


4.4

SPALDING & CHI METHOD

Similar to the above two methods in the Spalding-Chi


method, the compressible skin friction is given by the
incompressible form with appropriate correction factors to account
for compressibility and viscosity effects.

Aerodynamic Characteristics of Vehicle Components

CF =

1
Re F

CF

xe
Re x
inc

Fc

(4.9)

where

Fc =

T /T 1
aw
e

sin 1 A + sin 1 B

and

F Re x F c =

e
w

(4.10)

where expressions for A and B are as given in Eqn. 4.8.


One first computes Fc and then FRex. The equivalent
incompressible Reynolds number is given by Rexe FRex. Using this
equivalent Reynolds number, the corresponding incompressible
turbulent skin friction coefficient is determined using any one of
the well-known formulae such as Karman and Schoenherr,
(Eqn.4.5), Prandtl-Schlichting, (Eqn.4.6) or the Sivells & Payne
relation, viz.
CF

inc

0.088
log R

xe 1.5

Dividing the incompressible flow skin friction coefficient by


the factor Fc one obtains the desired compressible flow skin friction
coefficient.
For most flows, a portion of the flow over the body is
laminar. For these regions the mean friction coefficient is
determined from the approximate laminar flow value, viz.,

C F inc =

1.328
Re x

where the Reynolds number is based on the length from the leading
edge to the point of transition.
4.5

EMPIRICAL EQUATIONS

Over a wide range of angle of attack, velocity and altitude


conditions, Schmidt4,5 has presented data on skin friction
coefficients of hypersonic flow over flat plates both for laminar and
turbulent flows in terms of free stream parameters rather than the
flow parameters at the edge of the boundary layer. To this data

65

66

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

empirical curves have been fitted6 to obtain simpler relations


particularly useful in preliminary design. The suggested equations
are:
Laminar flow

C f lam

Re = 0.45 cos + 4.65

V
3050

sin cos 2.2

where, is the angle of attack and the V the velocity in m/s.


The above equation deviates no more than 20 per cent from
the data presented by Schmidt for low altitude, high angle of attack
flight and is closer to 10 per cent for the rest of the altitude and
angle of attack ranges:
15 45 ; 3050 V 8000 ; 45700

91400

Turbulent flow

turb

Re

0.2

V
= 0.048 sin 4.5 + 0.70
cos 2.25 sin 1.5

3050

(In the ranges of


5 50 ; 3050 V 8000 ; 30500 h 91500).

REFERENCES
1.

2.
3.

4.

5.

6.

Sommer, S.C. & Short, J.B. Free-flight measurements of


turbulent boundary layer skin friction in the presence of
severe aerodynamic heating at mach numbers from 2.8 to
7.0. NACA, March 1955. TN 3391.
Van Driest, E.R. Turbulent boundary layers In compressible
fluids. JAS, 1951, 18(3).
Spalding, D.B.M. & Chi, S.W. The drag of compressible
boundary layer on a smooth flat plate with and without heat
transfer. J. Fl. Mech., 1964, 18(1), Part 1.
Schmidt, J.F. Laminar skin friction and heat transfer
parameters for a flat plate at hypersonic speeds in terms of
free stream flow properties. NASA, September 1959. TN D-8.
Schmidt, J.F. Turbulent skin friction and heat transfer
parameters for a flat plate at hypersonic speeds in terms of
free stream flow properties. NASA, May 1961. TN D-869.
Hankey(Jr), W. L, & Alexander, G.L. Prediction of hypersonic
aerodynamic characteristics for lifting vehicles. September
1963. ASD-TDR-63-668.

CHAPTER 5
AERODYNAMIC HEATING AT HYPERSONIC
SPEEDS
5.1

INTRODUCTION

The rate of heat transfer to a vehicle due to aerodynamic


heating is a function of its geometry, orientation in space and its
flight trajectory. For flights at high supersonic and hypersonic
speeds, the problem of aerodynamic heating becomes important
and needs to be analysed particularly for Mach numbers greater
than about 2.0. The induced thermal stresses due to aerodynamic
heating can seriously affect the structural integrity of the vehicle
and may result in a component failure. Consequently, the
determination of the heating rates and the skin temperature are
needed to complement the pressure field calculation.
Several methods have been developed by different authors
to account for aerodynamic heating effects at high speeds. Some of
these are rigorous in which the full governing equations are
numerically solved on a computer for a given body geometry,
incidence and flight condition. These types of analyses are usually
reserved for complicated cases such as shock-boundary layer
interactions, flow separation, etc., or in final detailed design.
However, approximate analytical, semi-analytical and empirical
methods have been proposed for easy and fairly accurate and
acceptable calculations of heat transfer at high speeds that are
particularly useful in the preliminary analysis stage. No attempt is
being made here to collate all such methods and critically review
them. Instead some simple methods familiar to the author and
having good correlation with experiments or more exact analysis
are presented in this section.
5.2

HEATING ANALYSIS

The basic equation used to calculate the time rate of change


of heat stored in a thin skin element is given by

68

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

q& = ( w C p ,w w

d Tw
dt

= h (H st H w ) s F T w4

(5.1a)

for 3-dimensional stagnation points or unswept leading edges and

q& = ( w C p ,w w

dT w
dt

= h (H R H w ) s F T w4

(5.1b)

for swept leading edges, body, wing and control surfaces.


where
q

heating rate, J/m2-sec

w
Cp,w
w
Tw
t
h

density of surface material, kg/m3

specific heat of surface material, J/kg- K

wall or skin thickness, m

wall or skin temperature, K

time, sec

local heat transfer coefficient based on enthalpy,


kg/m2 sec

HR =

boundary layer recovery enthalpy, J/kg

Hw =

enthalpy at wall temperature condition, J/kg

emissivity of wall surface

Stefan-Boltzman constant, 5.67 10-8 W/m2-K - 4

radiation geometry factor normally taken as 1.0

Tw =

wall temperature, K

The above equation neglects the heat absorbed by the wall


from solar radiation and heat lost from internal radiation and
conduction. Normally, the radiation effects are compensating. For
cases in which the internal conduction is large, it could be
accounted for by incorporating an appropriate conduction term in
the above heat balance equation. It is to be noted that for high
supersonic and hypersonic flows, the heat transfer rate is expressed
in terms of the enthalpy whereas for temperatures below
approximately 800K the heat balance equation is normally
formulated in terms of temperatures.
As mentioned1 and to quote "to obtain good surface
temperatures and to a lesser extent good surface heating rates,
proper engineering judgement must be exercised to determine the
heat capacity in the above equation. Since the values of specific heat
of the surface material and the density of the surface material are

Aerodynamic Heating

thermal properties of the material, the only way to vary the heat
capacity significantly is to change the value of the material thickness.
For metallic surfaces, the thickness of the skin will give satisfactory
results. For surfaces that are insulated with low conductivity
insulation (such as space shuttle), a material thickness should be
used that will result in a heat capacity of approximately
2044 J/m2 - K".
The heat transfer rate varies from a maximum value at Tw = 0
(cold wall) to zero at Tw = Taw (adiabatic wall). Knowing the heat
transfer coefficient, the Mach number, angle of attack and altitude
combination, it is possible to compute the actual heating rate for
any wall temperature. From this, the time rate of change of skin
temperature can be determined if one knows the properties of the
wall material. By carrying out the heat transfer calculations at a
series of selected points along the vehicle trajectory, the time history
of the surface temperature of the body can be obtained.
The skin equilibrium temperature for each skin element
under consideration can be calculated from the above equation
when the time rate of change of temperature goes to zero.

4
h H R H eq = s F T eq

The temperature and heat flux for each skin element at time
J are calculated from

h (H R H j 1 ) s F T j41
T j = T j 1 +
t j
w c p , w w

and

q& j = h (H R H j ) s F T j4
respectively.
In practice, computational intervals are spaced closely
during the early times and farther apart at later times so that
precision is maintained during periods of rapid temperature rise
but excessive computational time will not be required after the
period of initial temperature rise.
5.3

STAGNATION POINT HEAT TRANSFER

5.3.1 Spherical Nose


The nose tip of a vehicle is usually blunt and spherical in
shape. Consequently a normal detached shock is present in front of

69

70

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

the stagnation point. From the known free stream conditions the
downstream properties of the shock are determined using the
normal shock relations for the case of perfect gas. If however, the
temperature is such that one has to consider the gas as real, then
the downstream properties of the shock are determined for an
equilibrium flow either by the use of real gas flow charts if one is
available or by numerical iterative solution as outlined in Andersons
book2. From the known properties downstream of the normal shock,
the stagnation enthalpy, recovery enthalpy and stagnation pressure
are determined as follows:
V 22
(for both perfect and real
2
gases). Recovery enthalpy, H R = H 2 + ( H 0 H 2 ) where, is the
recovery factor (In the above relations subscript 2 refers to
conditions downstream of the normal shock)

Stagnation enthalpy, H 0 = H 2 +

= Pr = for laminar flow, and

= (Pr )

for turbulent flow;

Pr = Prandtl Number.
The flow downstream of a normal shock at high speeds is
low subsonic. Hence, one can use the incompressible Bernoullis
equation without much of an error to calculate the stagnation
pressure both for perfect and real gas flows.

po

= p2 +

2 V 22
2

For a perfect gas the above value could be checked if


necessary with the exact calculation of stagnation pressure using
the isentropic relation as the downstream Mach number is known.
For real gases the condition of constant entropy can be used to
calculate the stagnation pressure by trial and error. However, these
are not really necessary as the downstream Mach number will be
very low subsonic at high Mach numbers and the incompressible
Bernoullis equation is more than adequate.
The commonly used expression to calculate the heat
transfer coefficient for a spherical nose stagnation point as given in
the book by White3 is

Aerodynamic Heating

h = 0.763 Pr

0. 6

( o o )

0.5

w w

o o

0. 1

du e

dx

(5.2)

The above expression is derived from the work of Fay and


Riddell4 but restricting to the case of non-reacting gases. It is
equally applicable to both perfect and real gases. (due/dx)s is the
velocity gradient at the stagnation point. Subscript o refers to
stagnation conditions downstream of the detached normal shock in
front of the body and w to the wall conditions at the stagnation
point.
A relation identical to Eqn. 5.2 but without the term
0.1


w w
is also mentioned in the book of Anderson2. This term

o o

is of the order of unity and alters the heat transfer rate by about 10
per cent depending on the wall temperature conditions.
The stagnation point streamwise velocity gradient is given
by the Newtonian impact theory, viz.,

du e
dx

2( p o p )

rN

where, rN is the spherical nose tip radius.


Supposedly a more accurate relation than that of the
Newtonian for the velocity gradient as given by Adams and
associates and mentioned by Dejarnette, et al 5 is
du

dx

=
r N

1.85

In the case of a perfect gas, a constant value of 0.7 is taken


for the Prandtl number and the viscosity is based on the Sutherland
law corresponding to the stagnation temperature. For a real gas the
stagnation temperature is determined from the known values of
stagnation pressure and stagnation enthalpy by curve fits6,7. For
real gases, the Prandtl number and viscosity are functions of
stagnation temperature and stagnation density, viz.,

71

72

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Pr = Pr (To, o) and o= o (To, o)


Their values are obtained by curve fits6,7. As the flow is
laminar near the stagnation region, the recovery factor dependence
on the Prandtl number is taken to be that corresponding to the
laminar flow, viz., Pr when calculating the heating rate, viz., q .
5.3.2 Cylinder Normal to the Stream
The stagnation point heat transfer coefficient for a 2dimensional circular cylinder normal to the incoming flow is given
by the expression3 viz.,

h = 0.57 Pr

0.6

0.5


w
w

o o

0.1

du

dx

(5.3)

The method of determining the relevant parameters in the


above expression is exactly identical to the axisymmetric case
described above. Comparison of Eqn. 5.3 above with the heat transfer
coefficient expression for the case of axisymmetric stagnation point
case, Eqn. 5.2 shows that both the expressions are the same except
for the leading numerical coefficient. All other terms being the same,
it is apparent that the stagnation point heating to a sphere is larger
than to a 2-dimensional cylinder.
5.3.3 Swept Wing Stagnation Line Heat Transfer
The wing or control surface leading edge is considered to be
cylindrical and flow past it is analogous to a flow past a yawed
inclined circular cylinder of infinite span. The detached shock will
be parallel to the leading edge whereas it is oblique to the incoming
free stream. Let be the sweep angle. The incoming velocity is
now decomposed in to components, V n = V C os , normal to the
leading edge and V t =V Sin , tangential to the leading edge. The
normal component of the incoming velocity is now made to pass
through the detached shock which is normal to it. The tangential
component of the velocity remains unchanged as it passes through
the shock. Let the resulting conditions downstream of the shock be
as

Aerodynamic Heating

p2

= pressure downstream of the normal shock

T2

= temperature downstream

= density downstream

V2 n = normal component of the velocity downstream


V2 t = tangential component of the velocity downstream
equal to V S in
S2

= downstream entropy or change in entropy

H2

= downstream enthalpy

All of the above quantities can be obtained from normal


shock tables for perfect gases and by an iterative numerical solution
of the governing equations across the shock for real gases2. After
passing through the shock, as the flow approaches the leading edge,
the normal component of the velocity viz., V2 n goes to zero on the
surface edge, leaving the tangential component V2 t unchanged,
giving rise to a non-zero velocity along the attachment line called
the stagnation line analogous to the stagnation point on a cylinder
normal to the free stream. The loss of the normal component of the
velocity on the surface results in increase of pressure and enthalpy
as follows:
Pressure on the surface can be calculated by the Bernoullis
equation as V2n is low subsonic
pe = p

V 22n

This value holds good both for perfect and real gases as
V2n is low subsonic.
5.3.4 Perfect Gas
For a perfect gas, the temperature on the surface is given by

1 2

Te = T 2 1 +
M 2n
2

where

2n

V2n

RT 2

73

74

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

and the density

e = 2

M 22n

1
1

The Mach number of the flow with a velocity V2t along the
stagnation line is
M

V 2t

RT e

The stagnation pressure, temperature, density and enthalpy


are

po = pe

M
1+
2

To = Te 1+
M
2

o = e

1+

H o = C p Te +

1
2

2
e

2
e

2
e

2
V 2t

respectively.
5.3.5 Real Gas
The enthalpy of the flow along the edge is given by

He = H2 +

V 22n
2

and the stagnation enthalpy is

Aerodynamic Heating

Ho = He +

2
V 2t

The stagnation pressure has to be obtained to know the


stagnation temperature. Since the entropy or the change in entropy
is known downstream of the shock, this value has to be constant as
the flow is decelerated isentropically from downstream of the shock
to the stagnation conditions. The stagnation pressure can be
determined from the Mollier diagram for equilibrium air (if one is
available) corresponding to the stagnation enthalpy and entropy.
Otherwise a trial and error iterative technique is adopted in which
the invariance of entropy is maintained. Once the stagnation
pressure is thus determined with the known value of stagnation
enthalpy, the other thermodynamic variables like temperature,
viscosity, etc., can be obtained by curve fits6,7.
5.3.6 Heat Transfer Coefficient h
Beckwith and Gallagher8 have suggested that the leading
edge heat transfer rate for a yawed cylinder is the same as that of a
normal cylinder but modified by a cosine factor involving the sweep
angle . Accordingly, the laminar flow stagnation line heat transfer
coefficient is

h = 0.57 Pr 0.6 o o

du e
dx

0.5

w w

o o

0.1

(cos )1.1

(5.4)

and the convective heat transfer rate is

l . e ,lam

= h ( H aw H w

At high Reynolds numbers and angles of sweep from 40 to


60, it was found that the boundary layer on a swept cylinder was
completely turbulent even at the stagnation line.
For turbulent flow the corresponding equation is

q i .e , turb = 1.04 Pr

0.6

(* * ) 0 . 8 ( V
( o ) 0. 6

sin
2

0.6

du e

dx

0.2

(5.5)

75

76

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

The superscript* denotes evaluation at the Eckert reference


enthalpy9 which is given by
H * = 0.5 ( H w + H e ) + 0.22 ( H aw H e

where
He = enthalpy at the outer edge of the boundary layer
For a perfect gas a reference temperature is used as follows
which is multiplied by Cp to get the reference enthalpy.

= 0 .5 ( T w + T e

The streamwise
streamline is given by

1
=

r le

sl

du

dx

) + 0.22 ( Taw Te )
velocity

gradient

at

the

stagnation

2 p e p

where, rle is the leading edge radius.


The nature of the flow, whether it is laminar or turbulent is
based on the free stream value of the Reynolds number based on
the leading edge diameter, viz.,
Re D =

V D

If the Reynolds number is below a specified lower limit value


it is laminar and, if it is above a specified upper limit it is turbulent.
In between these lower and upper limits the flow is considered
transitional. The heating rate in the transitional region is given by

q& = q&

lam

Re
Re low
D

+
Re upp Re low

q&
lam

Aerodynamic Heating

5.3.7 Heat Transfer on Flat Surfaces & Fuselage Panels


The approximate engineering method of determining the
heat transfer rate consists of treating the surface element under
consideration as a part of a flat plate. It has been shown by many
investigators that for a wide range of Mach numbers and
temperatures, a close approximation to the actual compressible
skin friction coefficient is obtained when the incompressible value
of the skin friction coefficient is evaluated at a temperature
corresponding to a reference enthalpy (or a reference temperature
for a perfect gas). Application of suitable Reynolds analogy factor
relating the skin friction coefficient to Stanton number is used to
get the appropriate heat transfer coefficient.
The basic expression for the heat transfer coefficient is the
same as given by Eqn. 5.1, viz.,

h =

w C p ,w w

dT w

+ s F T w4
dt
( HR Hw )

The recovery enthalpy is computed from

HR =H +

V 2 V e2
2

V e2
2

where, is the recovery factor.


For heat transfer determination it is necessary to know the
properties of the local flow parameters over the surface element
under consideration. These are obtained by any one of the
aerodynamic predictive methodologies such as shock expansion,
second order shock expansion, tangent cone, tangent wedge, etc.
For a real gas, knowledge of local pressure and entropy or enthalpy
by any of the approximate analysis methods is able to give all other
flow parameters.
The heat transfer coefficient h is given by

h =

*
f

Pr *
2

0.667

*Ve

77

78

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

where, Cf* is the skin friction coefficient and Pr* is the Prandtl
number
C

*
f

0.332

Re *

N lam

1
2

for a laminar flow, and


C *f =

0.185

Re *
log
10 N

tur

2.584

for a turbulent flow


The * conditions are evaluated at the reference enthalpy H *
for a real gas given by
H * = He + 0.5 (Hw He ) + 0.22 ( HR He )

(5.6)

H * = He + 0.5 (HR He )

(5.7)

or

known as adiabatic reference enthalpy


For a perfect gas for * conditions the reference temperature
T * is used given by
T * = 0.5 (Tw + Te ) + 0.22 (TR Te )

(5.8)

The subscripts e refer to the local conditions in the inviscid


shear layer or at the outer edge of the boundary layer, w the wall
conditions and, R adiabatic wall or boundary layer recovery
conditions respectively.
The Reynolds number is based on boundary layer running
length s and is given by
Re * =

*Ve s

and the Prandtl number

Aerodynamic Heating

Pr * =

*C
k

*
p

* = f (H *, pe ), for a real gas6,7

* = pe /R T* for a perfect gas.


T*

= f ( H *, pe ) for a real gas6,7

T*

= reference temperature as given by Eqn 5.8 above


for a perfect gas.

The viscosity is evaluated by using the Sutherland relation


at a temperature corresponding to T * and the thermal conductivity
corresponding to temperature T * and pe.
Substituting the above values of the skin friction coefficient
in the heat transfer coefficient relations we get
h H = 0.332 ( Pr * ) 0.667

*V e
Re *
N lam

(5.9)

for laminar flow, and

h H = 0.185 ( Pr * )

0.667

* V e

Re *
log10

N tur

2.584

(5.10)

for turbulent flow.


In the above expressions, Nlam and Ntur are the laminar and
turbulent Mangler transformation factors.

shown

Nlam

3 for a body of revolution and 1 for a planar body


(wing)

Ntur

2 for a body of revolution and 1 for a planar body


(wing)

Some experiments conducted on an X-15 aircraft10, have


that the correlation between the heat transfer

79

80

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

measurements and above prediction is better when the * conditions


were evaluated at the adiabatic reference enthalpy as defined in
Eqn. 5.7 above, instead of the reference enthalpy as proposed by
Eckert, Eqn 5.6.
5.4

HEAT TRANSFER ANALYSIS BY THE METHOD OF


QUINN & GONG

A real time aerodynamic heating programme for hypersonic


flight simulation has been developed by Quinn and Gong1. This
algorithm is capable of calculating 2- and 3-dimensional stagnation
point heating rates and surface temperatures. The leading edge
sweep is also accounted for in the program. In addition, upper and
lower surface heating rates and surface temperatures on flat plates,
wedges and cones can be calculated both for laminar and turbulent
flows together with boundary layer transition which is made a
function of free-stream Reynolds number and free-stream Mach
number. The results of this method of analysis when compared with
more exact values obtained by the use of a NASA in-house
aeroheating program showed that the heating rates and surface
temperatures as predicted by the real time heating analysis were
well within the required accuracy to evaluate heating trajectories.
One unique feature of this methodology is the use of free stream
conditions instead of the local flow quantities to calculate the heat
transfer rate. Because of its simplicity and good prediction
capabilities this method is described in detail.
5.4.1 Stagnation Point Heating Rate
The basic equation used to compute the surface
temperatures and heating rates (J/m2-s) for stagnation point
calculation is

q& = w C

p ,w

dT w
dt

= h H st H

FT 4

(5.11)

for 3-dimensional stagnation point or unswept leading edges.


For swept leading edges and surfaces the corresponding
expressions is

q& = w C

p ,w

dT w
dt

= h H R H

FT 4

(5.12)

The only difference between the above two equations is the


replacement of stagnation enthalpy by the recovery enthalpy for
swept leading edges and surfaces.

Aerodynamic Heating

To solve the above two equations we have to know the


heat transfer coefficient h. For this, a modified version of the
solution given by Fay and Riddell4 is used.
For 3-dimensional flow

h = 0.94 K 1 ( st st

) 0.4 ( w w )0.1 (dU / dx )x = 0

(5.13)

and for 2-dimensional flow

h = 0.706 K 2 ( st st

)0.4 ( w w )0.1 (dU / dx )x = 0

(5.14)

The subscripts st and w are stagnation and wall conditions.


K1 and K2 are 3-dimensional and 2-dimensional stagnation
factors respectively.
Quinn and Gong1 have used U.S. Customary Units in their
work. However, in the present work, S.I. Units are employed and so
whereever it is necessary, the relations taken from the work of
Quinn and Gong have been modified.
To solve the Eqns. 5.11- 5.14 suitably one has to know the
local flow conditions behind a normal shock wave. However, to
minimize the computational time, a method has been developed in
which the free stream values are used along with a table of values
for K1 and K2 maintaining the accuracy of the results.
The Eqns. 5.11- 5.14 are solved as
The stagnation enthalpy H (J/kg ) is given by

H st = H +

cos 2
(5.15)

for 3-dimensional calculations, and

HR =H+

+ 0.85

sin 2
2

(5.16)

81

82

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

for 2-dimensional calculations.

dU
1

=
dx
R

x=0

7 M

cos 2 1

cos 2

cos 2 + 5
P

(5.17)
where
R = body nose radius

st st =

T o K

= 2.43 10 7 T w o K

= 2.43 10

6 M 2 cos 2

M 2 cos 2 + 5

T
st

0.75

(5.18)

0.75

(5.19)

0.75

(5.20)

P w P st P

w =

Pw
287 T w

7M

cos
6

(5.21)

(5.22)

For 3-dimensional stagnation point calculations, the sweep


angle in the above equations will be zero. Values of H , Hw , TR

Aerodynamic Heating

and Tst are obtained by the curve fits given by Gupta et. al.6 and,
Srinivasn7. Values of K1 and K2 are obtained by linear interpolation
from the following table.
Table 5.1 Stagnation point heating factors
Mach No.

K1

K2

1.00

1.00

1.16

1.20

10

1.14

1.18

15

1.16

1.16

20

1.23

1.16

25

1.40

1.25

30

1.45

1.26

5.4.2 Convective Heating Equation for Small or Zero


Pressure Gradient Surfaces
The basic equation is the same as Eqn. 5.12 viz.,

dT
q& = (w C p ,w ) w = h (H R H w ) Tw4
dt
To determine the heat transfer coefficient h one has to know
the local flow conditions. However, for simplicity and to save time
the free stream values are utilized similar to the case of stagnation
point heat transfer analysis. The heat transfer coefficient is written
as

h= C 5

h +
o

(5.23)

where, ho is the heat transfer coefficient for a flat plate at zero angle
of attack and h is that portion of the heat transfer coefficient
caused by angle of attack and wedge or cone angles. For laminar
flow the suggested value of C5 is 1.73 and for turbulent flow C5 is
1.15.
For turbulent flow, the equations to calculate the heat
transfer coefficient are as follows:

83

84

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Lower Surface

ho

C 1 ( 0.0375 ) U

=
0.2
x

0.8
U

= A1

x 0.2

0.8

0.2

T

*
T

0.65

(5.24)

(5.25)

and for Upper surface

0.8
U

= A1

x 0.2

(5.26)

The constant C1 used is an empirical value used to account


for the approximation involved in developing the Eqn. 5.24. The
suggested best values for C1 are 1.0 for the upper surface and 0.90
for the lower surface. The difference between the heat transfer
coefficient calculated using the free stream flow conditions and the
local flow conditions are represented by the Eqns. 5.25 and 5.26.
Similarly for laminar flow, the relevant equations are
Lower Surface

ho

U

= 0.4
x

h = A2

0.5

T

*
T

0.125

(5.27)

0. 5

(5.28)

Aerodynamic Heating

Upper Surface

ho

U

= 0.421
x

h = A2

0.5

T

*
T

0.125

(5.29)

0. 5

(5.30)

In the above expressions, is the cone or wedge half angle


in degrees and the angle of attack in degrees. Values for A1 and A2
are as given in the tables to follow.
Table 5.2 Turbulent flow correction factors, A 1 10 4

Wedge or cone angle angle of attack, deg.


-1
0
+1
+5
+10

-10

-5

+20

+40

2
3
5

0.844
1.50
1.30

0.920
1.62
1.79

1.04
1.80
2.38

1.04
1.62
1.84

1.30
1.80
2.07

1.41
2.08
2.34

1.62
2.08
2.61

1.73
2.37
2.87

1.84
2.49
2.90

10

1.19

1.66

2.49

2.16

2.61

3.32

4.06

4.80

4.88

15

1.01

1.57

2.60

2.16

2.77

3.97

5.18

6.58

6.77

20
25

0.747
0.582

1.41
0.989

2.30
2.03

2.16
1.95

2.87
2.87

4.47
4.54

6.08
6.49

9.94
9.20

9.49
8.66

30

0.472

0.826

1.88

1.84

2.87

4.87

6.49

6.49

6.49

Table 5.3 Laminar flow correction factors, A 2 10 4


M

-10

-5

Wedge or cone angle angle of attack, deg.


-1
0
+1
+5
+10

+20

+40

0.366

0.378

0.390

0.354

0.366

0.378

0.403

0.415

0.427

0.488

0.507

0.533

0.488

0.528

0.547

0.561

0.558

0.561

0.732

0.799

0.871

0.763

0.455

0.950

1.01

1.04

0.819

10

0.997

1.18

1.51

1.46

1.62

1.85

2.03

2.07

1.87

15

1.13

1.53

2.20

1.83

2.29

2.68

2.94

2.96

2.51

20

1.17

1.77

2.48

2.51

2.90

3.54

3.82

3.75

3.20

25

1.12

1.74

2.75

2.48

3.33

4.15

4.27

3.66

3.11

30

1.03

1.71

2.95

2.30

3.66

4.27

4.27

2.44

2.44

85

86

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

It is to be noted that in Eqns. 5.25 and 5.28 the value for


wedge angle plus the angle of attack ( + ) is limited to 40, and in
Eqns. 5.26 and 5.30 the wedge angle minus the angle of attack
( - ) is limited to 10.
The values of the other parameters in the heat transfer
coefficient equations are

Re =

U x

= 2.43 10

H R = H + 0.85

0.75

For laminar flow,

H R = H + 0.89

and for turbulent flow,


H * = H + 0 .5

Hw = f
H

( H w H ) + 0 .22 ( H R H )

( Tw , P )

= f T , P

= f H

, P

In the above the values of Hw , H and T * are obtained from


curve fittings given6-7 and the values of A1 and A2 from the tables
above.
5.4.3 Boundary Layer Transition
The methods outlined above, utilizes only the free stream
flow properties to calculate the local heat transfer rates. In keeping

Aerodynamic Heating

up with the same methodology only the free stream Reynolds


number, free stream Mach number are utilized as criteria to
determine the boundary layer transition by Quinn and Gong as
follows:

[log Re

If log Re

[log Re

If log Re >

( )] the flow is assumed laminar

+Cm M

( )] the flow is assumed turbulent

+Cm M

Values of log R e T and Cm depend on type of flow, angle of


attack, sweep angle and leading edge or nose bluntness. The
authors suggest the following table of values, which according to
them produce satisfactory transition results.
Table 5.4 Transition Reynolds numbers and transition Mach
number coefficients.

(a) Conical flow


, deg

Cm
Sharp leading edge

log ReT

0-7
7 - 20
20 - 40

5.3
5.3
5.3

Blunt leading edge

0.25
0.20
0.15

0.20
0.18
0.12

(b) 2-Dimensional flow


Cm
,deg

Log ReT

Sharp leading edge


7 7< < 20

20

Blunt leading edge


7

7 < < 20

20

0 - 45

5.3

0.23

0.20

0.18

0.20

0.18

0.15

45 - 60

5.3

0.20

0.18

0.15

0.18

0.15

0.12

60 - 75

5.3

0.17

0.17

0.13

0.15

0.13

0.11

5.5

HIGH SPEED CONVECTIVE


METHODOLOGY OF TAUBER

HEAT

TRANSFER

An engineering type formulation to calculate the convective


heat transfer rates at high speeds has been made by Tauber11. The
final pertinent relations taken from that reference are as follows:

87

88

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

5.5.1 Stagnation Point Heat Transfer


The expression for the stagnation point heat transfer for a
sphere is

q& ws = 1.83 10 4



rn

0.5

Hw

V 3 1
Hs

watts/m2

(5.31)

where the nose radius rn is in metre, the free-stream density in


kg/m3, and the flight velocity V , in m/sec. Hw is the wall enthalpy
2
and Hs the stagnation enthalpy = H + V in J/kg.
2
5.5.2 Swept Infinite Cylinder
The analogous expression for the swept infinite cylinder is

q& w ,cyl = 1.29 10

r
cyl

Hw

V 3 1
H aw

0.5

(1 0.18 sin 2 eff )

cos eff

(5.32)

where

H aw = H + 0.5 V 2 1 0.18 sin 2 eff

sineff = sin cos ,

= the sweepback angle


In deriving the above equations the same Newtonian value
for velocity gradient at the stagnation point was used both for the
sphere and cylinder, viz.,

du

dx

2 p st p

1
=

rn
st

0.5

Aerodynamic Heating

As mentioned by Tauber in his report, the velocity gradient


calculated as above is correct for the case of a cylinder but a bit low
for the sphere. It has been suggested by him that if a correction is
made by using the appropriate value, the constant in Eqn. 5.31
changes from 1.83 to 1.90.
5.5.3 Cone and Flat Plate Heating Rates
5.5.3.1 Laminar boundary layer
The expressions for laminar boundary layer heating of
bodies without pressure gradients such as sharp cones, wedges and
flat plates at angle of attack is given by

cos c
q& w , Cone = 4.03 10 5
x

Hw

V 3.2 sin c 1
H aw

cos FP
q& w ,FP = 2.42 10 5
x

Hw

V 3.2 sin FP 1
H aw

0.5

W / m 2

(5.34)

0.5

W / m 2

(5.35)

where
Hw

= enthalpy of the wall

Haw = adiabatic wall enthalpy = H + 0.40V 2

= cone half angle

FP

= wedge angle or angle of attack of the flat plate

5.5.3.2 Turbulent Flow


For the case of flat plate and for V > 3960 m/sec

89

90

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

q& w ,FP

2. 45 10

)( sin 2 cos 2 . 62 )0 . 8
( x x bt )0.2

Hw

V 3.7 0.9
He

for 3960 > V > 1560 m /sec

qW FP =

) ( sin
(x x ) (T

3.72 10 4

cos 2.2

0 .2

bt

W /555

0 .8

0 .25

W /m 2

and for a cone

q&

w cone

= 1.15 q& w FP

where
Hw =

static enthalpy at the wall

He =

stagnation enthalpy at the outer edge of


boundary layer

distance from the leading edge for flat plate and


from nose for cone

xbt =

5.6

distance at which boundary layer transition


begins

EMPIRICAL EQUATION FOR CONVECTIVE HEAT


TRANSFER

Anderson2 in his book, reference has given a very simple


method to estimate aerodynamic heating at hypersonic speeds. The
heat transfer rate is expressed in a generalized form, viz.,

q& w = N V M C
The units for q w, and V are W/m2, kg/m3 and m/sec
respectively. The values of M, N and C are as follows:

Aerodynamic Heating

5.6.1 Stagnation Point


M = 3, N = 0.5

H
C = 1.83 10 8 R 1/2 1 w
Ho

where, R is the nose radius in metres, and H w and Ho are wall and
total enthalpies respectively.
5.6.2 Flat Plate in Laminar Flow
M = 3.2 and N = 0.5
C = 2.53 10

cos

0.5

sin x

0.5

Hw

1
Ho

where, is the local body angle with respect to the free stream and
x is the distance measured along the body surface in metres.
5.6.3 Flat Plate in Turbulent Flow
For V 3962 m/s, N = 0.8, M = 3.37 and

C = 3.89 10 8 (cos )1.78 (sin )1.6 (x T


Hw

1 1.11
Ho

For V > 3962 m/s,

Tw

556

)0.2

0.25

N = 0.8, M = 3.7 and

C = 2.2 10 9 (cos )2.08 (sin )1.6 (x T

) 0.2 1 1.11

Hw

H o

where, xT is the distance measured in metres along the body surface


in the turbulent boundary layer.
As mentioned by Anderson the above are useful for
preliminary analysis and are not recommended for more detailed
work .

91

92

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

REFERENCES
1.

2.
3.
4.
5.

6.

7.

8.

9.

10.

11.

Quinn, R.D. & Gong, L. Real-time aerodynamic heating and


surface temperature calculations for hypersonic flight
simulation. NASA, August, 1990. 4222.
Anderson(Jr), J.D. Hypersonic and high temperature gas
dynamics. McGraw-Hill Book Co. 1989.
White, F.M. Viscous fluid flow. McGraw Hill Book Co. 1991.
Fay, J.A. & Riddell, F.R. Theory of stagnation point heat
transfer in dissociated air. J. of Aero. Sci., 1958, 25(2), 73-85.
DeJarnette, F.R.; Hamilton, H.H.; Weilmuenster, K.J. &
Cheatwood, F.M. A review of some approximate methods in
aerodynamic heating analysis. J. Thermophysics, 1987, 1(1).
Gupta, R.N.; Lee, K.P.; Thompson, R.A. & Yos, J.M.
Calculations and curve fits of thermodynamic and transport
properties for equilibrium air to 30,000 K. NASA, October
1991. 1260.
Srinivasan, S.; Tannehill, J. & Weilmuenster, K.J. Simplified
curve fits for the thermodynamic properties of equilibrium
air. NASA , 1987. 1181.
Beckwith, I.E. & Gallagher, J.J. Local heat transfer and
recovery temperatures on a yawed cylinder at mach numbers
of 4.15 and high reynolds numbers. NASA, 1961. R-104.
Eckert, E.R.G. Survey of boundary layer heat transfer at high
velocities and high temperatures. Wright Air Development
Center, April 1960. 59-624 p.
Quinn, R.D. & Palitz, M. Comparison of measured and
calculated turbulent heat transfer on the X-15 airplane at
angles of attack up to 19.0. NASA, TM-X-1291.
Tauber, M.E. A review of high speed, convective, heat transfer
computational methods. NASA , July 1989. 2914.

PART - II
VALIDATION OF PREDICTION METHODS

CHAPTER 6
VALIDATION OF PREDICTION METHODS
Approximate methods are normally used in preliminary
design stages to predict flight control forces and moments
experienced by a flying vehicle in the entire speed range covering
the flight envelope. To ascertain the validity and the range of
applicability of these methods, particularly for hypersonic vehicles,
some detailed studies have been reported in the literature both in
the past and fairly recently. In general, the approach is to examine
several vehicle designs, such as the wing-body, the blended body,
the cone body, etc., which cover a broad range of proposed
hypersonic vehicle configurations and compare the predicted values
with the available experimental data and then draw conclusions.
In the subsonic and supersonic speed ranges, many
methods are available to predict the aerodynamic characteristics.
Some of these have been used for the analysis of a few hypersonic
vehicle configurations. The vortex lattice method for subsonic flow
analysis described by Lamar and Gloss1, which includes the leading
edge suction effects based on theory, described by Polhamus2 has
been applied to three different hypersonic vehicle configurations
and compared with the wind tunnel results3 at a Mach Number of
0.2. The theoretically predicted lift, drag due to lift and the pitching
moment, correlated well with the experimental results. Similar
results using the same type of analysis were also reported4. Since
the method described by Lamar and Gloss1 did not have the
capability to include vertical surfaces, it was not possible to predict
the lateral-directional characteristics of the configurations
analysed. For this reason, the vortex lattice method used was not of
much use in the preliminary design analysis. The most commonly
used prediction method in subsonic and supersonic flows is the
Panel method. Panel method has been applied to some specific
hypersonic vehicle configurations such as X-15, Space Shuttle and
Hypersonic Research Airplane, in the subsonic and supersonic

96

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

speed ranges. The results of these investigations are reviewed in the


later part of this work.
In supersonic speed ranges, the linear supersonic theory and
the hypersonic impact methods were applied to three non-slender
Hypersonic Airplane concepts at Mach numbers from 1.1 to 2.865. The
configurations used were similar to those of Penland, et al 3. Lift,
pitching moment, and drag due-to-lift values were calculated by the
planar method6. The mean camber surface of the body and the wing
were inputs to the linear theory program, but the vertical tail surfaces
that could contribute to the pitching moment were ignored. A drag
buildup analysis was carried out in which the drag due to skin friction,
the wave drag at zero angle of attack due to volume and the camber
drag at zero lift, were independently evaluated and summed up. The
hypersonic impact methods were also used to check the limit of
applicability of hypersonic methodology in the supersonic speed range.
Tangent-cone empirical method was used on the body and tangent
wedge method on the wing and tail surfaces. Prandtl-Meyer flow was
assumed on expansion surfaces, where the minimum expansion
pressure coefficient was limited to 1/M 2 . Also evaluated was the
tangent-cone empirical method on all the configuration components.
According to this work, the linear theory gave good lift prediction and
fair to good pitching-moment prediction over the Mach number ranges
tested, except in the transonic region. This good agreement between
data and the linear theory predictions of CL and Cm showed that the
linear theory with its thin-airfoil and planar assumptions was valid for
the class of low fineness ratio, blunt-base high volume vehicle
configuration. The linear theory drag prediction was generally poor
with good agreement only below M 1.2. The inaccuracy of the zero
lift drag prediction using the linear theory was attributed to the
violation of slender body assumption, as the vehicle configurations
had low-fineness ratio bodies. The tangent cone theory predictions
were good for lift, and fair to good for the pitching moment for M 2.0.
The combined tangent cone/tangent wedge theory (tangent cone for
the fuselage and tangent wedge for the wing and tail) gave the least
accurate prediction of lift and pitching moment. For all theories, the
zero lift drag was overestimated especially for M < 2. The tangent cone
method predicted the zero lift drag most accurately for M 2.0. The
level of agreement depended on the configuration being studied, very
good for one configuration but fair for others. No definite
conclusions were reached regarding the lower Mach number limit
of applicability of the hypersonic impact methodology.

Methods

An experimental investigation was conducted to provide a


systematic set of longitudinal characteristics and lateraldirectional stability data for a simplified wing-body model with a
series of vertical-tail arrangements7. The range of Mach numbers
covered were 1.60 to 2.86 at nominal angles of attack from 8 to
12. Comparisons were made of the experimental data at zero
angle of attack with three theoretical methods, a Second-order
Shock-expansion and Panel method (MISLIFT)8, a slender body and
first order panel method (APAS)22 and PAN AIR9, a higher order
panel method. Since the results from MISLIFT and APAS were
invariant with angle of attack, only the results from PAN AIR were
used for comparisons of the stability parameters at angles of
attack. Overall, the results were quite good except for a couple of
vertical tail configurations. The differences were attributed
probably to the inability of the program to account for the presence
of vortices, etc.
The conclusion reached were:
(a)

(b)

(c)

PAN AIR generally provided accurate predictions at moderate


angles of attack for complete configurations with either single
or twin vertical tails.
APAS provided fairly accurate predictions at zero angle of
attack for complete configurations with either single or twin
vertical tails.
MISLIFT provided estimates only for the simplest bodyvertical-tail configurations at zero angle of attack.

Experimental and theoretical aerodynamic characteristics


of two hypersonic cruise aircraft concepts at Mach numbers of 2.96,
3.96 and 4.63 were studied10. The test models consisted of a
fuselage with lenticular cross-section, two geometrically similar
wings, one set of horizontal tails sized for each wing, a wedge-centre
vertical tail, a set of fuselage mounted twin vertical tails, a flowthrough body-mounted nacelle, and a set of flow-through wingmounted nacelles formed other parts of the configurations. The
large wing airfoil had a circular arc upper surface and a flat lower
surface, whereas the smaller wing airfoil and the horizontal tail
airfoils were symmetric circular arcs. The airfoil of the twin vertical
tails was flat on the outboard surface and circular arc on the
inboard surface. Estimates from first order supersonic linear theory
and hypersonic impact theory were compared with the experimental
data. Two types of calculations were made under the hypersonic
impact method, viz.,

97

98

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

(a)

(b)

application of the tangent cone empirical calculations to all


the components in the impact region and Prandtl-Meyer
relations in the expansion region, and
the application of tangent cone theory to the fuselage and
tangent wedge theory to the wing, horizontal tail and vertical
tail with Prandtl-Meyer theory in the expansion regions.

In both the calculations, the minimum expansion was


limited to 0.7 vacuum pressure coefficient. To account for viscous
effects, Spalding-Chi method was utilised. For the linear theory
analysis, an integrated program11, which included an empirical skin
friction calculation based on the work of Sommer and Short12, was
used. The conclusions reached were:
(a)

(b)
(c)

The results of tangent-cone empirical theory to the fuselage


and tangent-wedge theory to the wing, horizontal and vertical
tails gave very good overall agreement with the experimental
data but the estimates of the aerodynamics of the individual
components were significantly different from the data.
The tangent-cone theory applied to all the configurations
generally showed poor estimates.
The predictions of first order supersonic theory were also bad
except for lift and drag data at low angles of attack.

A simple and rapid method for the computation of the


aerodynamic coefficients in the high Mach number range for small
angles of attack and tail deflection and arbitrary roll angles was
suggested13. The method was limited to the following configurations:
cylindrical bodies of circular cross-section with different nose
shapes, sharp cones, sharp ogives and hemispheres, arbitrary
position of the wing of any planform, consisting of flat plates with
sharp leading and trailing edges. The prediction methodology used
was the modified Newtonian, exact relations for oblique shocks, and
Prandtl-Meyer theory for expansion regions. Test results at M = 3.08
and M = 4.63 were compared with predictions and it was found that
the agreement was very satisfactory.
The aerodynamic characteristics of three hypersonic
configurations at a Mach number of 6 were studied3. Tangent cone
theory on the body, tangent wedge on the wing and vertical tail
surfaces and Prandtl-Meyer expansion for all expansion regions,
were utilised to compare the predicted data with experimental
values. The general trend of lift, pitching moment and drag were
observed but the comparison between the predicted drag and
pitching moment with experimental values were not too good.

7.67
C.G.OF MODEL

80

47.55
53.52
48.92

HORIZONTAL CONTROL
PIVOT

(a) BHVCB

31.60

12.52
19.25

Figure 6.1. HYFAC model (all dimensions in centimetres)

CANARD PIVOT

.810
0
STATION

6.5

MOMENT CENTRE

48.11

Methods
99

1.18
76.5

77.8

65

4.44

3.38

CANARDS

0.0238
RAD

0.665

3.46

4 36'

1 36'

HORIZONTAL CONTROL SURFACE

6.05

0.0254 RAD

0.292

8.95
3.485

3.92

70

2.10

(b) Components

65

6.3

Figure 6.1. HYFAC model (all dimensions in centimetres) (Concluded)

VERTICAL TAIL

1.382

1.49

0.47

8 35'

100
Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Methods

In the hypersonic speed range, the most commonly used


prediction method is the Gentry Hypersonic Arbitrary-Body
Aerodynamic Computer Program (HABP)14. Although this program
is more than thirty years old, it is still being used by many, with
some additions and modifications.
An experimental program was conducted at Mach 6 to
determine the aerodynamic characteristics of an all-body, delta
planform, hypersonic research aircraft (HYFAC Configuration)15.
The sketch of the model tested is given in Fig. 6.1. The body had a
delta planform with a 6.5 half-angle conical nose faired to an 80
swept leading edge afterbody. Aft of the conical nose the fuselage
had modified rhombic cross-sections. Computed theoretical values
were compared with experimental data. Predictions from various
methods like tangent cone, tangent wedge, shock-expansion,
Dahlem-Buck empirical, etc., were evaluated and compared with
experimental data. For skin friction effects, the reference
temperature method for laminar region and Spalding-Chi method
for turbulent layers were used. It was found that the tangent cone
for all the components gave the best correlation. For small control
deflections the agreement between predicted and experimental
values was good. However, for large control deflections, particularly
for negative deflections, none of the methods adequately predicted
the longitudinal characteristics. This was attributed to the use of
free stream dynamic pressure values over the control surfaces
rather than the local dynamic pressure. Some representative
comparisons between the predictions and wind tunnel
measurement taken from Clark15 is shown in Figs. 6.2, 6.3 & 6.4.
In these figures, the symbols, B, H, and V represent the body,
horizontal tail and vertical tail respectively. In Fig. 6.4, the side
force parameter CY/, (CY = side force coefficient), the effectivedihedral parameter Cl /, (Cl = the rolling moment coefficient),
and the directional stability parameter, Cn/, (Cn = yawing
moment coefficient) are respectively indicated by the symbols C Y ,
C 1 and C n . is the angle of side slip.

In order to determine the effect of wing leading-edge sweep


and wing translation on the aerodynamic characteristics of a wing
body configuration, a series of experiments were conducted at a Mach
Number of 616. Seven wings with leading edge sweep angles from 30
to 60 were tested on a common body over an angle of attack ranging
from 12 to 10 and side slip angles of 0 and 2. All wings had a

101

102

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Tangent cone on body and vertical tails andTangent cone on horizontal controls
Modified Newtonian (K = 2.4) on
horizontal controls
Shock expansion on horizontal controls
Tangent wedge on horizontal controls
Wind-tunnel data
h'deg
0
10
-30

0.04
0.03
0.02
0.01

h'deg

Cm

-30

0
-0.01

0
10

-0.02
-0.03
Tangent cone
Modified Newtonian (K = 2.4)
Shock expansion, and
Tangent wedge

4
3
2
1
L/D 0
-1

h'deg
- 30

-2
-3
-4
-0.15

-0.10

-0.05

0.05

0.10

0.15

0.20

0.25

CL

Figure 6.2. Comparison of several hypersonic theories with wind 106


tunnel data for HYFAC configuration (BHV) at R,
= 10.5
,l
(Continued).

Methods

Tangent cone on body and vertical tails andTangent cone on horizontal controls
Modified Newtonian (K = 2.4) on
horizontal controls
Shock expansion on horizontal controls
Tangent wedge on horizontal controls

.11
.10

h'deg
-30
0

Wind-tunnel data

.09

h'deg

.08
.07

0
10
-30

CD .06
.05
.04
.03
.02
.01
0

-30

20

h'deg
0

10

15

,
deg

10

-5
-.15

-.10

-.05

.05

.10

.15

20

.25

CL

Figure 6.2. Comparison of several hypersonic theories with windtunnel data for HYFAC configuration (BHV) at
R,l= 10.5 106 (Concluded).

103

104

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

.11

.10
.09
.08

h'deg
-30

.07
CD

.06
.05

-20

.04
.03

-10

.02
.01
0
Tangent cone on body, horizontal controls and
vertical tails
Wind-tunnel data

20

h'deg
10
0
-10
-20
-30

15

10
,deg

0
h'deg -30 -20 -10 0 10

-5
-.15

-.10

-.05

.05

.10

.15

.20

.25

CL

Figure 6.3. Comparison of tangent-cone theory with wind-tunnel


data for longitudinal aerodynamic characteristics of
HYFAC configuration BHV at R,l = 10.5 106 (Contd...).

Methods

.04
.03

h'deg
-30

.02
-20

Cm

.01
-5
0
5
10

0
-.10
-.02

Tangent cone on body, horizontal controls


and vertical tails
Wind-tunnel data

h'deg
10
5
0
-5
-10
-20
-30

4
3
2
1
L/D

0
-1
-2

h'deg
-30

-3
-4
-.15

-20
-10

-.10

-.05

.05

.10

.15

.20

.25

CL

Figure 6.3.

Comparison of tangent-cone theory with wind-tunnel data


for longitudinal aerodynamic characteristics of HYFAC
configuration BHV at R,l = 10.5 106 (Concluded).

105

106

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

CY

-.005

-.010
-.015
Wind-tunnel data (M = 6)

.001

Cl

-.001
-.002
Tangent cone on body, and horizontal controls

and -

Tangent cone on vertical tails


Tangent wedge on vertical tails
Modified Newtonian (K = 2.4) on vertical tails

.003
.002

Cn

.001
0
-.001
-4

12

16

20

, deg

Figure 6.4.

Comparison of several hypersonic theories with windtunnel data for lateral-directional stability characteristics
of HYFAC configuration (BHV) h= 0; R= 10.5 106.

Methods

common span, aspect ratio, taper ratio, planform area and thickness
ratio. The wings were translated longitudinally on the body to make
tests possible with the total and exposed mean aerodynamic chords
located at a fixed body station. The theoretical estimates were based
on Gentrys Program14. Tangent cone pressure distribution was applied
on the body and tangent wedge on the wings, (method 1). In the
expansion region a limiting expansion pressure coefficient of 70 per
cent of vacuum conditions, (i.e., Cp, limit= ( 1/M 2 ) was utilised for all
calculations. The base pressure on the body base was assumed to be
equal to the freestream static pressure. Spalding-Chi method was
used for viscous effects, assuming 100 per cent turbulent boundary
layer. The drag contribution from the body nose bluntness and the
wing leading and trailing edges were not taken into account as they
were estimated to be very low. An alternate analysis in which tangent
cone only on the fuselage fore body and tangent wedge on the wing
and body aft of the wing was also done (method 2). Comparison of the
wind tunnel data with the theoretical predictions lead to the following
conclusions:
(a)

(b)

(c)
(d)

Good to excellent predictions of longitudinal forces (normal,


axial, lift and drag) and lift to drag ratio were obtained
throughout the angle of attack range on all wing body
configurations tested by using the tangent cone theory on the
body and tangent wing theory on the wings and by limiting the
expansion pressure to 70 per cent vacuum conditions.
Unsatisfactory predictions of the magnitude of the pitchingmoment coefficients were obtained on all the wing-body
combinations tested with the tangent cone/tangent wedge
analysis. This was attributed to the wing body interference,
wing tip losses and the change in dynamic pressure over part
of the wing due to bow shock, etc.
The Gentry program was found unsatisfactory for estimating
the lateral-directional stability.
It was also suggested that the mean aerodynamic chord of the
exposed wing should be used as the design reference. The
independence of the longitudinal and lateral forces and
pitching moments from wing leading edge variations when the
exposed wing mean aerodynamic chords were located at a fixed
body station verified the Hypersonic Isolation Principle.

A wing cone configuration was identified as one of the


potential designs for a trans-atmospheric vehicle and a study was
conducted on a generic model17. The base line wing-cone model
consisted of a 5 half angle cone body, a cylindrical mid body, and a
9 truncated cone afterbody. The fuselage was fitted with a

107

108

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

4 per cent thick diamond airfoil delta wing of aspect ratio 1 that
could be located at three longitudinal positions while maintaining a
smooth wing body juncture. The model components included two
nose geometries which varied in bluntness, two canards which
differed in planform and three vertical tail arrangements. The test
Mach Number ranged from 2.5 to 4.5. The model was designed to
allow the wing to be positioned at five incidence angles ( 5, 2.5,
0, 2.5, and 5). Angle of attack was varied from 4 to 28 and
angle of slide slip from 8 to 8. Theoretical analysis was performed
using three prediction methods: The Gentry Hypersonic Arbitrary
Body Program14, Linear Theory18, and Supersonic Implicit Marching
Program (SIMP)19. The HABP method employed the tangent cone
theory for the body compression pressure and the tangent wedge
methodology for the wing, canard and tail compression pressures.
The Prandtl-Meyer expansion was used for lee-side pressures.
Spalding-Chi method was used for viscous forces. The linear theory
method was based on linearized supersonic potential theory and
slender body estimates for inviscid lift, far field wave drag using
supersonic area rule for inviscid zero lift drag and Sommer-Short
skin friction estimate12. The SIMP method solved the full potential
equation to provide inviscid characteristics and the skin friction
estimate was done using the Sommer-Short estimates.
Comparisons of the experimental data with several analysis
methods to predict the longitudinal aerodynamic characteristics of
the wing-body showed the following:

The HABP predicted lift and L/D values were overpredicted,


and the drag slightly underpredicted.

The stability level was reasonably well predicted, although


the non-linear aspects of the pitching moment curve were not
predicted.

Linear theory underpredicted lift and L/D, and overpredicted


drag and stability level.

The SIMP predictions agreed well with the experimental lift,


drag and pitching moment results although the absolute
pitching moment values differed.
HABP

results

for

the

wing-body

and

two

canard

configurations showed minimal changes in C L , CD,o and (L/D)max


due to the addition of the canard, consistent with experiments but
failed to predict the change in the centre-of-pressure location. For
the configuration with canard, Linear theory predictions showed a

Methods

small increase in CL, no change in CD,o and an increase in (L/D)max


compared to the baseline wing body combination. These results
were in contrast to the experimental data which showed no change
in CL, CD,o and (L/D)max. These discrepancies between linear theory
and experiment were attributed, perhaps to the inability of the
linear theory to accurately model the downwash of the canard and
its influence on the wing.
Three hypersonic vehicles, viz., Shuttle Orbiter, the FDl-7,
and the X-24C-10D were chosen20, to establish a rationale for
choosing the best combination of hypersonic analysis methods. In
this work, three Mach number ranges were defined: a low
hypersonic Mach number range from 3.0 6.5, a high hypersonic
Mach number range above 8.5, and a transitional region in
between. In the high hypersonic case, Modified-Newtonian was
suggested for all the components and the Prandtl-Meyer expansion
for the shadow regions. In the low hypersonic range, the methods
suggested were: Dahlem-Buck empirical for a round nose, tangentwedge for flat nose, inclined cone for a round body; tangent cone for
a strake: and tangent wedge for lifting surfaces. Different methods
for calculating viscous effects had very little impact on the overall
results and the Reference Temperature/Spalding-Chi method was
recommended. In the transitional hypersonic region, it was felt that
either one of the above low or high hypersonic approaches would
work well.
Gentrys program (HABP)14, has been in wide use since the
early 1970s to predict the hypersonic aerodynamic characteristics
of several vehicle configurations and their comparison with
experimental data as discussed above. However, it was felt by
Maughmer, et al 21, that there was a need for a comprehensive and
systematic study to explore the ability of the simple local surface
inclination methods to predict control forces and moments
generated by aerodynamic flight controls for a variety of
configurations in hypersonic flows. For their work they made use of
the industry developed single analysis program for subsonic,
supersonic and hypersonic flow regimes called the Aerodynamic
Preliminary Analysis II (APAS II)22.
The APAS II, is an aerodynamic analysis program based on
potential theory at subsonic/supersonic speeds and impact type
finite element solutions at hypersonic speeds. It was developed at
Rockwell International by integrating their version of the Woodward
subsonic/supersonic panel method, called the Unified Distributed
Panel (UDP) with an enhanced version of Gentrys program. The

109

110

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

first order panel method is capable of analyzing a complete aircraft


configuration with relatively short computation time. The method is
based on potential flow theory but includes surface leading edge
and side forces and semi empirical techniques for the calculation of
skin friction drag.
In addition to exploring the validity of hypersonic capability
of APAS methods, Maughmer et al, made use of the subsonic and
supersonic panel methods of that program, including wetted area
drag prediction and compared the results with experimental data in
all the speed ranges. In their work three representative hypersonic
vehicle configurations were considered, viz., (a) the X-15 Research
Aircraft, (b) the Hypersonic Research Airplane, (HRA), and (c) the
Space Shuttle Orbiter. For these vehicles experimental results were
available in the entire speed range from subsonic to hypersonic, so
that comparisons could be made between the predicted values and
experimental data. Being of a comprehensive nature and being fairly
recent, their work is considered in some detail.
6.1

NORTH AMERICAN X-15 RESEARCH AIRCRAFT

The North American X-15 was developed in the late 1950's


and test flown in 1960's. It was designed to reach flight velocities of
about 2000 m/s and an altitude of about 76,000 m. Wind tunnel
tests were done on the X-15 model at subsonic, transonic,
supersonic and hypersonic Mach numbers.
The fuselage was basically cylindrical in shape with fairings
along both sides. The wing had an aspect ratio of 2.5, a quarter
chord sweep angle of 25, and was equipped with conventional
trailing edge flaps for use during landing. The horizontal tail had a
quarter chord sweep angle of 45 and a dihedral angle of -15. This
all-movable tail was deflected asymmetrically for roll control.
During the development of the North American X-15, the
original upper and lower vertical tails were found to be insufficient
in producing the required stability. These original surfaces
consisted of a large upper vertical and small lower vertical, each
having a diamond shaped airfoil. Both the upper and lower verticals
were all movable and the rear portion of each surface could be
deflected to form speed brakes. The final vertical tails had wedge
shaped airfoils and the lower vertical was only slightly smaller than
the upper vertical. On both the upper and lower surfaces, the inner
portion was fixed and contained speed brakes while the outer
portion was all movable. The lower movable portion was jettisoned

Methods
Table 6.1

X-15 Research
characteristics)

airplane

(airplane

geometric

Wing (extended to body centre line)


Area, (sq ft)

200

Aspect ratio

2.50

Taper ratio

0.20

Mean aerodynamic chord, (in.)

123.23

Sweep of leading edge, (deg)

36.75

Span, (ft)

22.36

Root chord, (in.)

178.89

Tip chord, (in.)

35.78

Dihedral angle, (deg)

Incidence angle, (deg)

Twist, (deg)

Airfoil section

NACA 66005(modified)

Fuselage station for 20 per cent mean aerodynamics


chord, (in.)

339.19

Wing station for 20 per cent mean aerodynamic


chord, (in.)

52.17

Flap area, (sq ft)

15.48

Flap travel, (deg)

40
Wing (exposed)

Area, (sq ft)

105

Aspect ratio

2.15

Taper ratio

0.27

Root chord, (in.)

131.95

Tip chord, (in.)

35.78
Horizontal tail (exposed)

Area, (sq ft)

51.76

Aspect ratio

2.81

Taper ratio

0.21

Mean aerodynamic chord, (in.)

60.07

Sweep of quarter-chord line, (deg)

45

Span, overall, (ft)

17.64

Root chord, (in.)

84.27

Tip chord, (in.)

25.28

Dihedral angle, (deg)

-15

111

112

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows


Airfoil section

NACCA66005(modified)

Fuselage: station for 50 per cent horizontal-tail mean


aerodynamic chord, (in.)
Span station for 50 per cent horizontal-tail mean
aerodynamic chord, from fuselage, (in.)
Tail arm, 20 per cent wing mean aerodynamic chord to
50 per cent horizontal-tail mean aerodynamic chord, (in.)
Incidence range, normal to plane of symmetry, (deg)
Pitch control
Roll control

537.52
26.96
1 98.33
35 (down),
15 (up)
+7.5

Vertical tail (upper, exposed)


Area, (sq ft)

40.8

Aspect ratio

1.03

Taper ratio

0.74

Mean aerodynamic chord, (in.)

1 07.5

Sweep of leading edge, (deg)

30

Span, (exposed), (in.)

55

Root chord, (in.)

1 22.5

Tip chord, (in.)

90.75

Airfoil section

10 wedge

Fuseulage station for 50 per cent vertical-tail


mean aerodynamics chord, (in.)
Span station for 50 per cent vertical-tail mean
aerodynamic chord, from fuselage, (in.)
Tail arm, 20 per cent wing mean aerodynamic chord to
50 per cent vertical-tail mean aerodynamic chord, (in.)

5 20.25
26.15
1 81.06

Movable outboard panel area, (sq ft)

26.5

Angular travel of movable area, (deg)

+ 7.5

Vertical tail (lower, exposed)


Area, (sq ft)

34.2

Aspect ratio

0.785

Taper ratio

0.79

Mean aerodynamic chord, (in.)

1 09.2

Sweep of leading edge, (deg)

30

Span, (exposed), (in.)

44

Root chord, (in.)

1 21.4

Tip chord, (in.)

96

Airfoil section

10 wedge

Fuseulage station for 50 per cent vertical-tail


mean aerodynamics chord, (in.)

519.4

Methods
Span station for 50 per cent vertical-tail mean
aerodynamic chord, from fuselage, (in.)
Tail arm, 20 per cent wing mean aerodynamic chord to
50 per cent vertical-tail mean aerodynamic chord, (in.)

21.15
1 80.21

Movable (jettisonable) area, (sq ft)

19.9

Angular travel of movable area, (deg)

+ 7.5
Fuselage

Length, high-speed nose, (ft)

49.17

Length, low-speed nose, less boom, (ft)

50.16

Width, including side fairings, station


346 to station 411, (in.)

88.0

Height, station 186 to station 530, (in.)

56.0

Maximum cross-sectional area, (sq ft)

21.4

Fineness ratio, average

9.4

Nose apex angle, (deg)

31.0
Speed Brakes (Upper and Lower)

Location hinge line, fuselage station, (in.)


Side area, each, (sq ft)
Angular travel, fuselage centre line, (deg)

534
4.88
41

113

114

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

50.16

17.64

22.36

Figure 6.5.

Three view drawing of the X-15 airplane (All dimensions


in feet).

Methods

for landing. The original tail configuration results were used21, for
comparisons in the subsonic and transonic speed ranges and the
final configuration results in the supersonic and hypersonic speeds.
The X-15 aircraft geometric characteristics are given in
Table 6.1. The three view drawing of the aircraft is given in Fig.
6.5. Since, this aircraft configuration was extensively studied in
1960's during its development, it is interesting to study the
comparisons
between
theoretical
predictions
and
measurements made then 23 (period in which the finite element
panel methods for subsonic and supersonic flows had not yet
been developed), and now fairly recently 21, wherein, there has
been an extensive use of combined subsonic/supersonic
computer codes available for analysis. A summary of these
studies is as follows:
6.1.1

Walker & Wolowiczs Work

The stability and control derivatives for the X-15 research


airplane in power-off flight at supersonic and hypersonic Mach
numbers have been presented in this work, both as derived from
the then existing theoretical methods and as measured in various
wind tunnel facilities. Calculations were made for Mach numbers
within and beyond the estimated flight envelope and for angles of
attack from 0 to 25. The results were compared with the
experimental data in the Mach number range from 2 to
approximately 7 and, for the static derivatives, with the limiting
values given by the Newtonian theory. Because the report23 is old,
originally classified and not easily available, some details of the
prediction methodology used in their work is presented below.
6.1.2

Lift Characteristics

The lift for the complete airplane is calculated by the


method24, in which the total lift is considered initially to be the sum
of the individual lifts of the exposed wing and horizontal-tail
surfaces and of the fuselage, each treated as an isolated body.
Incremental lifts are then added which represent corrections for the
interference that arises when the components are placed adjacent
to one another in the overall configuration. The interference is
reciprocal, consisting of reflection-plane and upwash effects on the
wing due to the presence of the fuselage, and of the carryover lift on
the fuselage due to the exposed wing and tail panels. Both these
effects were treated as wing contributions in accordance with the
method described by Pitts, et al 24. The forces on the horizontal-tail

115

116

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

surfaces at zero incidence (controls fixed) were similarly treated.


According to the procedure24, the X-15 lift coefficient can be
expressed as

CL =

SW
C LW (K W B + K B W )
S
S cos T
C LT (K T B + K B T
+Q T
S

) 1
+ C LB

(6.1)

where the K terms represent the interference factors which account


for the lift of the wing and the horizontal tail in the presence of the
body, KWB and KTB, and for the lift of the body in the presence of the
wing and the horizontal tail, KBW and KBT respectively. In the above
equation, the aerodynamic coefficients when used without a
superscript, were based on the dimensions of the wing with leading
and trailing edges extended to the plane of symmetry of the
airplane, and when primed, on the dimensions of the isolated
surface or body. Sw and ST are the exposed wing and tail areas
respectively, T is the dihedral angle of the horizontal tail measured
from the x-y plane, positive when rotated upwards, is the
downwash angle at the tail in degrees.

Q =

q 1 (C L )1
q (C L )

where, q is the free stream dynamic pressure and q1 is the dynamic


pressure of downstream flow where the local stream pressure
behind the bow shock wave is equal to the free stream static
pressure; (C L )1 is the lift curve slope in the downstream based on
the local Mach number and (C L ) is the lift curve slope in freestream flow.
6.1.3

Wing

Assuming that both the wing and the tail surfaces can be
considered as flat plates for the lift analysis, Van Dykes unified
small disturbance theory was used for calculating the lift
coefficient. This method is suitable to both the supersonic and
hypersonic speed regimes. For a 2-dimensional flat plate wing at an
angle of attack ,

Methods

c n = (C p )low er (C p )u p p er

cn = 2

+1
+

+1

4
H

(6.2)

1 1 H 1 1

where, H is the similarity parameter M

(6.3)

1 .

The lift as calculated above was converted from


2-dimensional to 3-dimensional lift at hypersonic speeds by the
following approximation based on the linear theory.
C L
C N' =

4
M

cn

(6.4)

In the above expression + L is the lift curve slope from linear


theory for the 3-dimensional planform as given in ref 25, and cn as
given by the Eqn. 6.3. Neglecting the wave drag and skin friction
drag, the lift coefficient for the isolated wing becomes

C L W = C N cos

(6.5)

while that for the wing in the presence of the body (based on
reference area S ),
C L
C L W = (K W B + K B W )

4
M

cn

SW
S

cos

(6.6)

Approximate values for the interference terms, KWB and KBW


taken from Pitts, et al 24, were substituted in the above equation to
get the value of CLW. The Newtonian theory limit was also evaluated
with KWB = 1 and KBW = 0.

117

118

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

6.1.4

Horizontal Tail

The lift characteristics of the horizontal tail at zero


incidence were calculated in a similar manner to that of the wing
with dihedral angle taken into account. The fuselage-induced
upwash at the tail plane was considered negligible and also the
term KTB corresponding to KWB in Eqn. 6.1 was taken as unity. The
wing downwash parameter d/d was estimated from the charts of
Haefeli, et al 26. This downwash was found to be negligible for Mach
number values greater than 4.0.
6.1.5

Fuselage

The lift from the fuselage was derived both from the inviscid
flow and the viscous crossflow. Second-order Shock-Expansion
theory was used to calculate the inviscid lift. Since the fuselage
cross-section of the X-15 aircraft was noncircular, the second-order
shock-expansion expressions, as given in Syvertsons report27, was
multiplied by a factor equal to the ratio of the actual planform area
to that of an equivalent body of revolution having the same local
cross-sectional area as the X-15 configuration. This approximation
lead to the relationship for the inviscid flow as

(C )
LB

In which

in viscid

(C )
'
N

= R

( )

SB
C N
S

cos

(6.7)

was obtained from Syvertson & Dennis27

(Appendix C), and


R =

T otal fuselage plan form area


Planform area of equ ivalen t body of revolu tion

The lift due to viscous crossflow was calculated using the


Allen and Perkins theory28

(C )
LB

viscou s

= c d c

A 2
cos
S

(6.8)

The term A, is the planform area of the fuselage, consisting


of forebody area only (vertex to the leading edge approximately,
since the wing and tail in effect block the crossflow over the
remaining sections). A value of c d = 1.2 was taken in the overall
c
Mach Number and angle of attack ranges.
Newtonian theory was applied approximately assuming that
the X-15 fuselage may be represented from the vertex to a station

Methods

immediately rearward of the canopy by a circular cone and over the


remaining length by a cylinder of constant diamond-shaped crosssection similar to that of the combined fuselage and side fairings.
6.1.6

Pitching-Moment Characteristics

From the calculated values of the lift coefficients for


various components as determined above and from the centre-ofpressure charts given in Pitts, et al24, the buildup of the moments
about a centre of gravity location of 20 per cent of the mean
aerodynamic chord (based on area S, the reference area equal to
area of wing with leading and trailing edges extended to the plane of
symmetry) was calculated as follows.
6.1.7

Wing & Horizontal Tail

The moment arm for the lift of the wing in the presence of
the body differs in general from that for the lift induced by the wing
on the body, with the difference depending primarily on Mach
number and fuselage diameter. The moments from the two sources
therefore must be determined separately. For consistency with the
lift calculations described earlier, both effects are charged to the
wing. The characteristics for the horizontal tail at zero incidence
are also determined likewise, although the moment arms for the
various interference effects, due to the absence of the fuselage
afterbody, are essentially equal. The pitching moment of the
combined wing and tail in the presence of the fuselage is given by

C m W +C m T =

SW
x
x

C LW K W B W B + K B W B W

S
c
c
S Tcos T
x
d
+ Q
C L T 1 (K TB + K B T ) T
d
S
c

(6.9)

where, x is the longitudinal distance from centre of gravity to centre


of pressure of component lift measured in direction of fuselage
centre line. The subscripts WB, BW, TB, and BT refer to wing in the
presence of the fuselage, fuselage in the presence of the wing,
horizontal tail in the presence of the fuselage, and fuselage in the
presence of horizontal tail respectively. c is mean aerodynamic
chord. Other symbols are defined in the earlier lift section.

119

.2

.4

.6

0
4
8
,deg

12

Horizontal tail off

16

Horizontal tail on
(iT=0)

.1

Airplane
trimmed

(a) M = 2.01

20

Wing

Horizontal tail on, wind tunnel (ref.1)


Calculated

0
-.1

Horizontal
tail off

-.2
Cm

-.3

Horizontal
tail on

-.4

Wing

-.5

Figure 6.6. Comparison of calculated and experimental lift and pitching-moment characteristics for the X-15
airplane at various Mach numbers.

CL

.8

1.0

1.2

1.4

120
Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

CL

.2

.4

.6

.8

1.0

1.2

1.4

Horizontal tail on

12
,deg

16

Wing

} wind tunnel (ref.3)

Horizontal tail off

Horizontal tail on,


Horizontal tail off
Calculated

(b) M = 2.29

.1

Figure 6.6. Continued

20

Airplane
trimmed

-.1

Horizontal
tail off

-.2

Cm

-.3

Wing

Horizontal tail
on

-.4

-.5

Methods
121

CL

.2

.4

.6

.8

1.0

1.2

1.4

12

.1

(c) M = 2.98

20

Airplane
trimmed

Figure 6.6. Continued

16

Wing

, deg

Horizontal tail off

Horizontal tail on

Horizontal tail on,


Horizontal tail off wind tunnel .3)
Calculated

0
-.1

Horizontal
tail off

-.2
Cm

-.3

Wing

Horizontal
tail on

-.4

-.5

122
Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

.2

.4

CL .6

.8

1.0

1.2

12
(d) M = 4.65

20

.1

Airplane
trimmed

Figure 6.6. Continued

16

,deg

Wing

Horizontal tail on

Horizontal tail off

Horizontal tail on, wind tunnel (ref.3)


Horizontal tail off
Calculated

-.1

Horizontal
tail off

-.2
Cm

Wing

-.3

Horizontal
tail on

-.4

Methods
123

CL

.2

.4

.6

.8

1.0

Horizontal tail off

12

Horizontal tail on

20

24

Figure 6.6. Concluded

(e) M = 6.86

,deg

Wing

16

Horizontal tail on,


Horizontal tail off wind tunnel (ref.1)
Calculated

.1

Airplane
trimmed

-.1

Cm

-.2

Wing

Horizontal
tail on

Horizontal
tail off

-.3

124
Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Methods

22

Calculated
(iT, deg (refs. 3,4)

20

20

15

18
16
14

, deg

12

10
8
6
4
2

iT = 15

iT = 15

iT =20

iT =20

0
(a) M = 2.29

(b) M = 2.98

20
18
16
14
12

, deg 10
8
6
4

iT = 15

iT = 20

iT = 20

iT = 15

2
0
-.4

-.2

0
Cm

.2

(c) M = 4.65

.4 -.4

-.2

0
Cm

.2

.4

(d) M = 6.86

Figure 6.7. Comparison of the calculated and experimental


stabilizer effectiveness of the X-15 airplane for
incidence settings of 150 (leading edge up) and -200
(leading edge down) at various Mach numbers.

125

126

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

20

M=2.98

M=2.29

M=4.65

M=6.86

16

, deg

12
8
4
0

-.008 -.004
0
C n , per deg
R

20

-.008

Cn

-.004
0
, per deg
R

-.008

-.004
0
, per deg

Cn

-.008

Cn

-.004
0
, per deg
R

Calculated
Data
(refs. 3,4)

16
12
, deg
8
4
0

M=2.98

M=2.29

-.001

C l

.001

, per deg
R

-.001

C l

.001

, per deg
R

M=6.86

M=4.65

-.001

.001

C l , per deg
R

-.001

C l

.001

, per deg
R

Figure 6.8. Comparison of the calculated and experimental directionalcontrol derivatives for the X-15 airplane at several Mach
numbers.

Methods

20

M=2.98

M=2.29

M=6.86

16

M=4.65

12
, deg
8
4

Calculated
Data
(refs. 3,4)
0

C l1'

.001 .002
, per deg

C l1'

-.001 .002
, per deg

20

C l1'

.001 .002
, per deg

-.001

C l1'

M=4.65

M=2.98

M=2.29

002 003
, per deg

M=6.86

16
12
, deg
8
4
0

.001 .002

C n1

'T '

, per deg

.001 .002

C n1

'T '

, per deg

.001 .002

C n 1'

, per deg

T'

.001 .002 .003

C n 1'

, per deg

T'

Figure 6.9. Comparison of the calculated and experimental lateralcontrol derivatives for the X-15 airplane at several
Mach numbers.

127

128

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

6.1.8

Fuselage

The centre of pressure for the lift due to the inviscid flow
over the fuselage was calculated by the Second-order ShockExpansion method27, (Appendix C) and that due to viscous cross
flow by the procedure described by Perkins & Jorgensen29. The
former was found to vary slightly with the Mach number and the
latter to be essentially constant. The moment coefficient for the
fuselage was expressed as
x
xB

+ C L
C m B = C L B B

c in viscid B c viscous

(6.10)

Figures 6.6 to 6.9 show some of the comparisons of the


experimental data with the theoretical predictions in the Mach
number ranges of 2.01 to 6.86. The figures are reproduced from
Walker and Wolowicz 23.
Walker also deals with the study of longitudinal control
characteristics such as lift variations due to incidence as well as
angle of attack for wing body combinations, damping in pitch and
lateral directional derivatives by the methods outlined in
Pitts et al 24.
The main conclusions reached were that the calculated
longitudinal characteristics for the most part were in close accord
with the results from the wind tunnel data. The lateral and
directional characteristics agreed well with the wind tunnel results
in the lower angle-of-attack range. However, due to the interference
of the bow shock wave on the tail surfaces and other effects not
accounted for in the theory, some disagreement was found at higher
angles for the stabilizer effectiveness and several of the lateral
directional characteristics at high angles of attack. Both stability
and controllability were maintained well beyond the estimated
speed limit.
Further, it was found that the limiting values predicted by
the Newtonian theory for the static derivatives were found, in
general, to be lower than the trends shown by the unified
supersonic-hypersonic small disturbance theory used for wing and
tail lift calculations and the second-order shock-expansion theory
used for flow over the fuselage.
6.1.9

Maughmer et al. Analysis of X-15

As mentioned earlier, the APAS code was used for the


analysis. Panel method was the basis for subsonic and supersonic

Methods

flows. For the hypersonic analysis, the impact methods chosen for
the various components were: tangent cone empirical for the bodies;
tangent wedge empirical for the surfaces; and modified Newtonian
with the factor K =C p m ax

for the leading edges and blunt trailing

edges and blunt end of the fuselage. Prandtl-Meyer empirical was


used for all shadow regions except for blunt ends or trailing edges
where the high Mach number base pressure method was used.
Viscous corrections were calculated assuming 100 per cent
turbulent boundary layer and the use of reference temperature/
Spalding-Chi method. The HABP had a choice of whether or not
to ignore the aerodynamic contributions of components
shielded(or shadowed) in the wake of upstream parts of the vehicle.
For the X-15 aircraft it was found that the shielding had very little
impact on the results and as such the shielding option was not
utilised.
Predicted results were compared with experimental values
at Mach numbers of 0.56, 0.8, 1.03, 1.18, 2.96, 4.65 and 6.83,
covering the entire speed range. Some representative comparison
figures taken from the above work are reproduced in Figs. 6.10 to
6.17. Summary of their findings are as follows:
At very low speeds, the lift and drag were well predicted up
to angle of attack of 25, both for zero flap deflection and 40 flap
deflection. The change in lift and drag due to elevator control
deflection were all reasonably well predicted although the absolute
values of the coefficients were somewhat in error. However, this
was not the case with the pitching moment versus angle of attack
and versus lift coefficient. The general trend of the changes in
pitching moment coefficient with angle of attack and CL with
elevator deflections were captured but not the values. Regarding
the lateral/directional coefficients, viz., the side force, yawing
moment and rolling moment, due to aileron and vertical tail
deflections, the predictions were good enough for conceptual design
purposes. The inclusion of edge forces did not make much
difference in the results.
In the transonic speed regime, the longitudinal
characteristics were compared with the panel method. The lift and
drag predictions at angles of attack up to 20 were good with and
without control deflections. Similar to the low speed case, the
pitching moment values were unsatisfactory apart from the correct
trend in the changes with horizontal tail deflection.

129

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

0.3

PITCHING MOMENT COEFFICIENT

NASA RM
L57D09 APAS II

0.2

f
40
0

0.1

0.0

-0.1

-0.2

10

40

30

20
ANGLE of ATTACK (deg)

0.3

0.2
PITCHING MOMENT COEFFICIENT

130

0.1

0.0

-0.1

-0.2

0.0

0.5

1.0

1.5

2.0

2.5

3.0

LIFT COEFFICIENT

Figure 6.10. Comparison of experimental and theoretical pitching


moment coefficients for the North American X-15 at Mach
0.056 for flap deflections of 00 and 400.

Methods

0.4
NASA RM
TM X-24 APAS II

PITCHING MOMENT COEFFICIENT

0.2

h
0
-3
-6

a
0
3
0

0.0

0.2

0.4
0.5

0.0

0.5

1.0

1.5

LIFT COEFFICIENT

Figure 6.11. Comparison of experimental and theoretical pitching


moment coefficients for the North American X-15 at Mach
1.18 for various horizontal tail deflections.

131

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

2.0
NASA RM
TM X-820 APAS II

DRAG COEFFICIENT

1.5

h
0
-20
-35
-45

1.0

0.5

0.0

-10

10

30

20

40

50

ANGLE OF ATTACK (deg)

2.0

1.5
DRAG COEFFICIENT

132

1.0

0.5

0.0

-0.5

0.0

0.5

1.0

1.5

2.0

LIFT COEFFICIENT

Figure 6.12. Comparison of experimental and theoretical drag


coefficients for the North American X-15 at Mach 2.96
for various horizontal tail deflections.

Methods

PITCHING MOMENT COEFFICIENT

0.5

0.0

0.5

1.0
0.5

NASA
TM X-820

0.0

APAS II

h
0
-20
-35
-45

0.5

1.0

1.5

2.0

LIFT COEFFICIENT

Figure 6.13. Comparison of experimental and theoretical pitching


moment coefficients for the North American X-15 at
Mach 2.96 for various horizontal tail deflections.

133

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

1.00

NASA TM X-236
APAS II
APAS II with Shielding

0.75

LIFT COEFFICIENT

0.50

0.25

0.00

0.25
5

15

10

20

25

ANGLE OF ATTACK (deg)

0.4

0.3
DRAG COEFFICIENT

134

0.2

0.1

0.0
0.2

0.0

0.2

0.4

0.6

0.8

LIFT COEFFICIENT

Figure 6.14. Comparison of experimental and theoretical longitudinal


aerodynamic coefficients for the North American X-15
at Mach 6.83 showing the effect of including the
hypersonic shielding option.

Methods

NACA RM L57D09
APAS II Potential + LE +Tip

CY

0.000

0.002

0.004

10

20

30

40

ANGLE OF ATTACK (deg)

Cn

0.002

0.000

0.002
0

10

20

30

40

30

40

ANGLE OF ATTACK (deg)

Cl

0.004

0.002

0.000

10

20
ANGLE OF ATTACK (deg)

Figure 6.15. Comparison of experimental and theoretical vertical tail


effectiveness for the North American X-15 at Mach
0.056.

135

136

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

NASA TM X-820
APAS II Potential
APAS II Potential + LE
APAS II Potential + LE + Tip

0.000

CY

0.005

0.010
10

10
20
ANGLE OF ATTACK (deg)

30

40

0.004

Cn

0.002

0.000
10

0
10
20
ANGLE OF ATTACK (deg)

30

40

0.001

Cl

0.000

0.001
10

10
20
ANGLE OF ATTACK (deg)

30

40

Figure 6.16. Comparison of experimental and theoretical vertical tail


effectiveness for the North American X-15 at Mach 2.96.

Methods

NASA TM X-236
APAS II

0.050

C Y

0.025

0.000

10
5
ANGLE OF ATTACK (deg)

15

20

15

20

15

20

0.000

Cn

0.002

0.004
5

10

ANGLE OF ATTACK (deg)


0.001

C l

0.000

0.001
5

10

ANGLE OF ATTACK (deg)

Figure 6.17. Comparison of experimental and theoretical vertical tail


effectiveness for the North American X-15 at Mach 6.83.

137

138

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

At supersonic speeds, the lift and drag predictions were


good. The prediction of the slope of moment coefficient against lift
curve was much better when compared with the results at low
speed. The prediction of lateral/directional coefficients missed
the nonlinear behavior observed in the experimental results but
was found acceptable for preliminary design work, provided the
angle of attack was not too great.
The comparison of theoretical and experimental results of
the longitudinal data for the hypersonic Mach numbers showed
that the lift, drag and the pitching moment coefficients were good.
Regarding the lateral/directional coefficients, at low angles of
attack, the comparison between theoretical values and
experimental data was reasonably good for the case of aileron
deflection. However, at higher angles of attack, only the sign of
the force and moment were predicted correctly. The predictions
of side force, yawing moment, and rolling moment due to vertical
tail deflection were relatively good and acceptable for conceptual
design studies.
Comparison of the work of Walker and Wolowicz 23 done in
1960 with that of Maugmer, et al.21, in 1991, on the same X-15
aircraft reveal the following: both at supersonic and low
hypersonic speeds the theoretical results obtained by using the
second-order shock-expansion method for the fuselage,
inclusion of viscous crossflow contribution as proposed by Allen
and Perkins, and the use of the small disturbance pressure
coefficients for hypersonic flows with the unified similarity
parameter as suggested by Van Dyke for lifting surfaces and
utilizing the data presented in the report by Pitts, et al.24 for the
interference effects between various components of the vehicle
and their method of determining the lift and centre of pressure of
wing-body tail combinations, etc., gave very good correlation with
the experimental data. However, in this work, the drag was not
computed. Prediction of the longitudinal characteristics and the
directional derivatives due to vertical tail deflections were as good
and somewhat better than that presented in Maughmer, et al.21,
where the panel method was utilised for analysis in the
supersonic case. In the hypersonic regime, the predictions from
both the Gentrys program contained in Bonner, et al.22 and the
prediction methods used in Walker and Wolowicz 23, agreed
equally well with the experimental data.

Methods

6.2

HYPERSONIC RESEARCH AIRPLANE

A wing-body concept for a Hypersonic Research Airplane


(HRA) was developed at NASA Langley Research Center in the mid
1970's to serve as a hypersonic flight technology demonstrator.
The wind tunnel results on this configuration are reported3, 30, 31, 32
in the Mach number ranges from subsonic to low hypersonic.
The configuration of the vehicle consisted of a body, a
cropped delta wing having a 2.1 negative incidence and 10
dihedral and a centre vertical tail. The airfoil was a modified
circular arc with a leading edge radius (normal to L.E.) followed by
a wedge section. It was equipped with full span elevons that could
be deflected symmetrically for pitch control and asymmetrically
for roll control. The centre vertical tail had a dual hinge line that
allowed for a diamond shaped airfoil at subsonic and supersonic
speeds. At hypersonic speeds, the rear portion deflected to form
a wedge shaped airfoil. They could also be deflected to form
speed brakes at high speeds. The dual hinged rudder could also
be deflected in the same direction to provide yaw control.
Dillon and Pittman32 deal with an experimental
investigation of the static aerodynamic characteristics of scale
model of the above research airplane at a Mach number 6 and
comparison of the measurements with theoretical predictions
using Gentrys method. The geometrical characteristics of the
model tested are given in Table 6.2, the model dimensions in
Fig. 6.18. Comparisons between some of the predicted values
and wind tunnel measurements, taken from the above referenced
report are reproduced in Figs. 6.19 to 6.21. In these figures, the
symbols B, BW, and BWVCH refer to the body or fuselage, the body
and wing, and the combination of body, wing and centre vertical
tail (wedge airfoil used in hypersonic testing) respectively. The
following options were used in the Gentrys program for
theoretical predictions: for compression regions, tangent cone on
the fuselage and tangent wedge on the wing and vertical tail; for
expansion regions, Prandtl-Meyer expansion; and for skin
friction, Spalding-Chi method (with 100 per cent turbulent
boundary layer). The conclusions reached were that, in general,
the Gentrys Hypersonic Arbitrary Body Aerodynamics computer
program gave reasonable predictions of longitudinal aerodynamic
characteristics. At lower angles of attack the lift was
underpredicted and drag and pitching moment overpredicted,
whereas, at the higher angles of attack the lift was overpredicted,
and the drag and pitching moment were underpredicted.
However, lateral-directional stability parameters were not well
predicted except for the isolated body.

139

140

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows


Table 6.2

Geometric characteristics of model (Hypersonic


Research Airplane)
Wing

Area (includes fuselage intercept), m2 (in2)

0.060 (92.63)

Area, exposed, m2 (in2)

0.030 (47.00)

Area, wetted, m2 (in2)

0.064 (98.98)

Span, m (in.)

0.244 (9.62)

Aspect ratio

0.999

Root chord (at fuselage center line), m (in.)

0.371 (14.59)

Tip Chord, m (in.)

0.119 (4.7)

Taper ratio

0.322

Mean aerodynamic chord


(includes fuselage intercept), m (in.)

0.294 (11.57)

Sweepback angles:
Leading edge, deg

67.5

25-perent chord line,deg

61.1

Trailing edge, deg


Dihedral

angle,deg

Incidence

10

angle,deg

-2.1

Airfoil thickness ratio:


Exposed root

0.051

Tip

0.078

Leading-edge radius (normal to leading edge),cm (in.)

0.064 (0.025)

Trailing-edge thickness, cm (in.)

0.064 (0.025)

Elevons:
Tip chord, percent wing tip

36.6

Span, percent total span

59.8

Area, both, m2 (in2)

0.0064 (9.89)

Vertical Tail
Area, exposed, m2 (in2)

0.007 (10.93)

Span, exposed m (in.)

0.077(3.06)

Aspect ratio of exposed area

0.857

Root chord at fuselage surface line, m (in.)

0.101 (3.99)

Tip chord, m (in.)

0.057 (2.256)

Taper ratio

0.565

Mean aerodynamic chord exposed area, m (in.)

0.097(3.804)

Sweepback angles:
Leading edge, deg

49.9

Trailong edge, deg

18.5

Methods

Vertical Tail
Hinge line location, per cent chord

68.7

Arudder/Atotal

0.295

Leading-edge, radius, cm (in.)

0.064 (0.025)

Fuselage
Length, m (in.)

0.584 (23.0)

Nose radius, cm (in.)

0.159 (0.063)

Maximum height, m (in.)

0.076 (2.98)

Maximum width, m (in.)

0.097 (3.83)

Fineness ratio of equivalent round body

6.86

Planform area, m2 (in2)

0.042(65.12)

Wetted area:
Without components of base, m2 (in2)

0.122 (188.6)

With wing on, m2 (in)

0.116 (179.4)

Ab, m2 (in2)

0.0023 (3.54)

Complete model:
Planform area, m2 (in2)

0.072 (112.12)

Aspect ratio of planform

0.825

141

10

.045
.132

x-STATION
0.0 .045

.018

.132
.408

.531
.561

7.6

.075

7.6

.075

(a) Model details

MODEL REFERENCE
LINE

.365

1.350

.297
.212
.182 .254

.026

air foil section at exposed wing root

.506

.204

10 .016 .238

.057

.687

67.5

.65

c.g.

AIRFOIL SECTION AT WING TIP

.751

.830
.870
HL

1.088

y-STATION
.209
1.000 1.075

HL

Y-STATION
.209

Figure 6. 18. Model general dimensions. All dimensions have been normalized by body length (l = 58.4 cm)

.254
.212
.182

.531
.408
.297

.561

.751

.830

.870

X-STATION
1.000

BASE PRESSURE
MEASUREMENT
LOCATION

142
Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Methods
7.75

12

.0312

12
.0226

12

.0022
.0161
.0673
.0981

HL
.1326

49.9
1.8

22.8
.0216
.1735
.1485

12
SUBSONIC AIRFOIL

12
.0384

HYPERSONIC AIRFOIL

7.75
12
.0454

SPEED BRAKES

(b) Vertical tail variations


Figure 6.18. Model general dimensions. All dimensions have been
normalized by body length (l = 58.4 cm) (Concluded).

143

144

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

.025

.020

.015

.010
Cm

.005

0
- .005

-.010

20

15
10
, deg

5
Exp.

Theory

.15

.20

BW
BWVCH

-5
-.10

-.05

.05

.10

.25

CL

Figure 6.19(a). Comparison of experiment with theory for body


buildup

.30

Methods

0
L/D

-2

-4

.12

.10

Exp.

Theory

B
BW

.08

BWVCH

.06
CD

.04

.02

0
-.10

-.05

.05

.15

.10

.2.0

.25

.30

CL

Figure 6.19(b). Comparison of experiment with theory for body


buildup (Concluded).

145

146

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows


Exp.

Theory

B
BW
BWVCH

CY

-.02

.002

Cl

-.002

C n -.002

-.004
-5

10

5
, deg

15

.20

Figure 6.20. Comparison of experiment with theory for effect of body


buildup on static lateral-directional stability
characteristics = 2
2.

Methods

.015
.010
.005
Cm

0
-.005
-.010
-.015
-.020

20
15
10
, deg

-5
e,deg Exp. Theory

10
5
0
-5
-10
-15

0
0
-5
-.10

-.05

.05

.10

.15

.20

.25

.30

.35

CL

Figure 6.21(a). Comparison of experimental with theory for effects


of elevon deflection on BWV CH configuration.

147

148

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

.18
e, deg Exp. Theory

10
5
0
-5
-10
-15

.16
.14
.12

.10
CD

.08

.06
.04

.02
0
0

0
0
0
0
-.10

-.05

.05

.10

.15

.20

.25

.30

.35

CL

Figure 6.21(b). Comparison of experimental with theory for effects


of elevon deflection on BWV CH configuration (Continued).

Methods

4
2
0
0
0
L/D

0
e, deg Exp. Theory

10
5
0
-5
-10
-15

0
-2
-4
-.10

-.05

.05

.10

.15

.20

.25

.30

.35

CL

Figure 6.21. Comparison of experimental with theory for effects of


elevon deflection on BWV CH configuration (Concluded).

The same aircraft design was analysed by Maughmer et.


al.21 using the APAS code at Mach numbers of 0.80, 0.98, 1.20
and 6.00. The main results of their work were as follows:
a)

The agreement between the panel method predictions with


experimental data both at the low and high speeds for the
longitudinal aerodynamic characteristics were good, as in
the case of X-15. The lift coefficient was slightly
underpredicted possibly due to the fact that the vortex lift
was not fully taken into account. The change in lift
coefficient with symmetric elevon deflection was predicted
well enough for conceptual design. The pitching moment
coefficient predictions were much better for HRA than for
X15.

b)

In the case of lateral/directional coefficients, the change in


side force, yawing moment, and rolling moment with

149

150

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

asynchronous elevon deflection as predicted by the


potential flow plus the leading edge suction force agreed
well the experimental data. However, the changes in the
coefficients due to vertical tail deflection were not at all
predicted well and were typically in error by 50 per cent or
more. Further, for these cases the trends in general were
also not captured.
c)

In the hypersonic flow, the predictions were computed


using HABP, both with and without shielding. For
longitudinal cases, the shielded values compared better
with the experimental data. For lateral coefficients with
aileron deflections, the predictions without shielding agreed
somewhat better with experimental data than those with
shielding. The lateral derivatives due to vertical tail
deflection at zero angle of attack agreed well at M = 6.0 only,
unlike the low and supersonic Mach number cases.
Figs. 6.22 to 6.26 illustrate some of the typical comparisons
between the predictions based on APAS code and wind
tunnel experiments as reported21.

Methods

PITCHING MOMENT COEFFICIENT

0.10

0.05

0.00

-0.05

-0.10
-5

10

15

25

20

ANGLE OF ATTACK (deg)

0.10

PITCHING MOMENT COEFFICIENT

NASA
TP-1189

APAS II

0
-5
-10
-15

0.05

0.00

-0.05

-0.10
-0.4

-0.2

0.0

0.2

0.4

0.6

0.8

LIFT COEFFICIENT

Figure 6.22. Comparison of experimental and theoretical pitching


moment coefficients for the hypersonic research airplane
at Mach 0.20 for various elevon deflections.

151

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

PITCHING MOMENT COEFFICIENT

0.02

0.01

0.00

NASA
TP-1249 APAS II

-0.01

e,

10
5

0
-5
-10
-15

-0.02

-0.03
-5

10

15

20

ANGLE OF ATTACK (deg)

0.02
PITCHING MOMENT COEFFICIENT

152

0.01

0.00

-0.01

-0.02

-0.03
-0.1

0.0

0.1

0.2

0.3

0.4

LIFT COEFFICIENT

Figure 6.23. Comparison of experimental and theoretical pitching


moment coefficients for the hypersonic research airplane
at Mach 6.0 for various elevon deflections. The
hypersonic shielding option was not included in the
theoretical calculations.

Methods

NASA TP-1189
APAS II POTENTIAL
APAS II POTENTAIL + LE
APAS II POTENTAIL + LE + TIP

0.004

0.002

0.000
-5

10

15

20

15

20

15

20

ANGLE OF ATTACK (deg)

Cn

0.000

-0.002

-0.004
-5

10

ANGLE OF ATTACK (deg)

Cl

0.001

0.000

-0.001
-5

10

ANGLE OF ATTACK (deg)

Figure 6.24. Comparison of experimental and theoretical rudder


effectiveness for the hypersonic research airplane at
Mach 0.20.

153

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

NASA TP-1249
APAS II
APAS II WITH SHIELDING

CY

0.001

0.000

-0.001
-5

10

15

20

15

20

15

20

ANGLE OF ATTACK (deg)

Cn

0.000

-0.001

-0.002
-5

10

ANGLE OF ATTACK (deg)

0.002

Cl

154

0.001

0.000
-5

10

ANGLE OF ATTACK (deg)

Figure 6.25. Comparison of experimental and theoretical rudder


effectiveness for the hypersonic research airplane at
Mach 6.0 showing the effect of including the hypersonic
shielding option in the theoretical calculations.

Methods

EXPER

APAS II

CY

0.004

0.002

0.000
0

MACH NUMBER

Cn

0.000

-0.002

-0.004
0

2
MACH NUMBER

Cl

0.001

0.000

-0.001
0

2
MACH NUMBER

Figure 6.26. Comparison of experimental and theoretical rudder


effectiveness as a function of Mach number for the
Hypersonic research airplane.

155

156

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

6.3

SPACE SHUTTLE ORBITER

The Space Shuttle Orbiter configuration consists of a


body, a double delta wing, a body flap and a centre vertical tail.
The double delta wing is equipped with full span elevons broken
into two panels in each side. These can be deflected
symmetrically as an elevator for longitudinal control or
asymmetrically as ailerons for roll control. The body flap, located
on the bottom side at the rear of the Orbiter is used as the primary
trim device. Body flap deflections range from +22.5 (trailing edge
down) to 11.7. The vertical tail consists of a split rudder that
can be deflected together for yaw control and separated to act as
a speed brake. The Space Shuttle Orbiter physical geometry and
hypersonic analysis model figure used in the APAS analysis is
given in Fig. 6.27.
Results of the comparison of the predicted values by the
application of APAS code with the experimental data for three
different Mach number values as reported by Maughmer, et. al.
are as follows:
*

Longitudinal aerodynamic coefficients as predicted by the


subsonic panel method of APAS code were compared with
the experimental data at a Mach number of 0.25. Included
in the comparisons were the results of full span elevon deflections from - 20 to +10 .
*
The predicted lift coefficients for the most part were within
10 per cent of the experimental data although the lift curve
slopes were not well predicted. However, the flap effectiveness of the elevon CL was predicted reasonably well except
at low angles of attack and large negative deflection angles.
*
Drag predictions were also good for preliminary design
analysis purposes.
*
The pitching moment coefficients as a function of angle of
attack did not agree well.
*
The slopes of the predicted and experimental curves were
sometimes of opposite sign.
*
Regarding the lateral control derivatives, the potential flow
modified with the leading edge suction analogy only yielded
the best agreement with experiments.
*
Similar were the cases for directional control derivatives.
The same kind of results were also observed with the tests
at Mach 0.8.
Comparison of the predictions using the HABP without
employing the shadowing of downstream components at the
hypersonic Mach number of 5.0 showed excellent agreement of

Figure 6. 27. Space shuttle orbiter physical geometry and hypersonic analysis model

Methods
157

158

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

the lift coefficient upto an angle of attack of 20 and remained


reasonably good till even upto 40. The predicted CD vs and CD
vs CL curves agreed well with the experiments. The flap
effectiveness for symmetrical elevon deflection was also well
predicted. The characteristic trends of the pitching moment
coefficient variation with angle of attack and lift were captured.
However, the control effectiveness was significantly overpredicted
for large negative elevon deflections. The changes in the values of
CD, CL and Cm due to the Shuttle body flap deflections as observed
in the experiments were more or less captured by the HABP. The
lift and drag data were not affected or influenced much whether
the shielding due to upstream influences were taken into
account or not. However, for the case of the pitching moment the
shielded results were better at low angles of attack while the
unshielded results were better at higher angles of attack.
Comparisons between HABP predictions without shielding for the
control derivatives showed that both the trend and magnitudes
were captured quite well. But this was not the case for the
directional control derivatives. Similar comparisons were also
exhibited for Mach 20 case. It was noted that HABP predictions
improve as the Mach number increases. Some representative
figures taken from Maughmers work showing comparisons
between theoretical predictions and experiments are reproduced
in Figs. 6.28 to 6.36.
CONCLUSIONS
From a critical study of the comparisons between various
prediction methods and the experimental data on a wide variety
of hypersonic vehicles, some of which have been described
above, the following general conclusions are drawn.
At subsonic speeds, the vortex lattice and the panel
methods give reasonably accurate values for the lift and drag in
the linear angle of attack range. But the pitching moment was not
well predicted. For the case of lateral and directional coefficients
the potential flow panel method with leading edge suction
analogy yields the best agreement with the experiments, and the
results are satisfactory for preliminary design analysis. The vortex
lattice method that was used to analyze some hypersonic
configurations in subsonic flow did not have the ability to include
vertical surfaces, and for this reason it could not be used to
determine lateral-directional control characteristics.

Methods

PITCHING MOMENT COEFFICIENT

0.4

0.2

0.0

-0.2
-5

10

20

15

ANGLE OF ATTACK (deg)

0.4

PITCHING MOMENT COEFFICIENT

ADDB

APAS II

e,

10
0

-10
-20

0.2

0.0

-0.2
-1.0

-0.5

0.0

0.5

1.0

1.5

LIFT COEFFICIENT

Figure 6.28. Comparison of experimental and theoretical pitching


moment coefficients for the space shuttle orbiter at
Mach 0.25 for various elevon deflections.

159

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

1.25
ADDB

DRAG COEFFICIENT

1.00

APAS II

10
0

-10
-20

0.75

0.50

0.25

0.00
-10

20

10

30

40

ANGLE OF ATTACK (deg)

1.25

1.00
DRAG COEFFICIENT

160

0.75

0.50

0.25

0.00
-0.25

0.00

0.25

0.50

0.75

1.00

1.25

LIFT COEFFICIENT

Figure 6.29. Comparison of experimental and theoretical drag


coefficients for the space shuttle orbitter at Mach 5.0
for various elevon deflections. The hypersonic shielding
option was not included in the theoretical calculations.

Methods

PITCHING MOMENT COEFFICIENT

0.05

0.00

-0.05
ADDB

APAS II

10
0

-0.10

-10
-20

-0.15

-10

10

20

30

40

ANGLE OF ATTACK (deg)

PITCHING MOMENT COEFFICIENT

0.05

0.00

-0.05

-0.10

-0.15
-0.25

0.00

0.25

0.50

0.75

1.00

1.25

LIFT COEFFICIENT

Figure 6.30. Comparison of experimental and theoretical pitching


moment coefficients for the space shuttle orbiter at
Mach 5.0 for various elevon deflections. The hypersonic
shielding option was not included in the theoretical
calculations.

161

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

1.5

LIFT COEFFICIENT

ADDB
APAS II
APAS II WITH SHIELDING

1.0

0.5

0.0

-0.5
0

-10

20

10

30

40

ANGLE OF ATTACK (deg)

1.00

DRAG COEFFICIENT

162

0.75

0.50

0.25

0.00
-0.25

0.00

0.25

0.50

0.75

1.00

1.25

LIFT COEFFICIENT

Figure 6.31. Comparison of experimental and theoretical longitudinal


aerodynamic coefficients for the space shuttle orbiter
at Mach 5.0 showing the effect of including the
hypersonic shielding option.

Methods

PITCHING MOMENT COEFFICIENT

1.00
ADDB

APAS II

10
0

0.75

-10
-20

0.50

0.25

0.00
-10

10

30

20

40

ANGLE OF ATTACK (deg)

PITCHING MOMENT COEFFICIENT

1.00

0.75

0.50

0.25

0.00
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

LIFT COEFFICIENT

Figure 6.32. Comparison of experimental and theoretical drag


coefficients for the space shuttle orbiter at Mach 20.0
for various elevon deflections. The hypersonic shielding
option was included in the theoretical calculations.

163

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

0.04

ADDB

APAS II

bf

-11.7
10.0
22.5

c L

0.02

0.00

-0.02
-10

10

20

30

40

30

40

ANGLE OF ATTACK (deg)

0.06

0.04

0.02

c D

164

0.00

-0.02
-10

10

20

ANGLE OF ATTACK (deg)

Figure 6.33. Comparison of experimental and theoretical increments


in longitudinal aerodynamic coefficients for the space
shuttle orbiter at Mach 20.0 for various body flap
deflections.

Methods

ADDB
SHUTTLE FLIGHT TEST
APAS II

CY

0.001

0.000

-0.001
-10

10

20

30

40

30

40

30

40

ANGLE OF ATTACK (deg)

Cn

0.001

0.000

-0.001
-10

10

20

ANGLE OF ATTACK (deg)

Cl

.002

.001

.000
-10

10

20

ANGLE OF ATTACK (deg)

Figure 6.34. Comparison of experimental and theoretical increments


in lateral force and moment coefficients due to
differential elevon deflections for the space shuttle
orbiter at Mach 5.0.

165

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

ADDB

0.002

APAS II

0
10

SHUTTLE FLIGHT TEST

CY

0.000

-0.002

-0.004
0

10

15

20

15

20

15

20

MACH NUMBER

X10-3

0.5

Cn

1.0

0.0
-0.5
0

10
MACH NUMBER

.0050

Cl

166

.0025

.0000
0

10
MACH NUMBER

Figure 6.35. Comparison of experimental and theoretical increments


in lateral force and moment coefficients due to
differential elevon deflections as a function of Mach
Number for the space shuttle orbiter.

Methods

ADDB
APAS II
APAS II WITH SHIELDING

CY

0.001

0.000

-0.001
-10

10

20

30

40

30

40

30

40

ANGLE OF ATTACK (deg)

0.000

Cn

0.001

-0.001
-10

10

20

ANGLE OF ATTACK (deg)

Cl

0.001

0.000

-0.001
-10

10

20

ANGLE OF ATTACK (deg)

Figure 6.36. Comparison of experimental and theoretical rudder


effectiveness for the space shuttle orbitter at Mach 20.0
showing the effect of including the hypersonic shielding
option in the theoretical calculations.

167

168

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

At low and moderate supersonic speeds, the PAN AIR


method predicts the longitudinal aerodynamic characteristics
very well. Due to the separated flow, some results at supersonic
speeds for the lateral/directional control derivatives are found to
be unacceptable. The rolling moment derivative due to aileron
deflection and the yawing moment derivative due to rudder
deflection are predicted well enough for conceptual design
purposes. The methods based on the linear supersonic flow
theory and slender body approximations give reasonably good
predictions only in the low supersonic speed ranges. Utilisation
of the second-order shock-expansion theory for the body, Van
Dykes unified hypersonic-supersonic similarity relations and the
approximate methods of the Pitts, Neilsen, and Kaattari24 for the
interference and downwash effects gave very good predictions in
the supersonic and as well as low hypersonic speed ranges, both
for longitudinal and lateral control aerodynamic characteristics.
In the hypersonic flow regimes there are no intermediate
theoretical models between the simple flow inclination methods
and the nonlinear CFD methods (Euler or Navier Stokes). Setting
up of computational grids as well as its execution requirements,
make CFD method unsuitable for preliminary design analysis.
For this reason, the original HABP as developed by Gentry or with
slight improvements or additions to the basic program as
contained in APAS is being universally used as a prediction
methodology at hypersonic speeds. To account for viscous
effects, the most commonly adopted method is Spalding-Chi. The
longitudinal aerodynamic characteristics particularly the lift and
drag are well predicted by the HABP when the tangent cone
theory is applied on the body for compressive forces and the
tangent wedge on the lifting surfaces and the Prandtl-Meyer
expansion for the lee-side surfaces with limiting value of a
pressure coefficient of 70 per cent of vacuum conditions. The
magnitudes of the pitching moment coefficients are not always
well predicted in almost all of the cases tested due to errors in the
locations of the centres of pressures. However, the trends in the
changes in the pitching moment as a function of angle of attack
and flap or elevon deflections, are generally captured. Lateral
control has been well predicted when the shielding is not taken
into account. The prediction methodology using the surface
inclination method improves as the Mach number increases.
Some conventional aircraft configurations like X-15 may be
influenced by body interference effects at low hypersonic Mach
numbers, which are not accounted for in the HABP. For cases

Methods

like this it is better to use in addition, an alternate approach like


the one suggested by Walker and Wolowicz23 that utilizes the
procedures outlined in the work of Pitts, et. al 24 to account for
interference and downwash effects between various components
of the vehicle.
REFERENCES
1.

2.

3.

4.

5.

6.

7.

8.

9.

Lamar, J.E. & Gloss, B.B. Subsonic aerodynamic


characteristics of interacting lifting surfaces with separated
flow around sharp edges predicted by a vortex lattice
method, NASA, 1975. TND-7921.
Polhamus, E.C. Concept of the vortex lift of sharp-edge
delta wings based on a leading-edge suction analogy, NASA,
1966. TND-3767.
Penland, J.A.; Dillon, J.L. & Pittman, J.L. An aerodynamic
analysis of several hypersonic research airplane concepts
from M=0.2 to 6.0. 16th Aerospace Sciences Meeting, AIAA.
January 1978. pp. 78-150.
Pittman, J.L. & Dillon, J.L. Vortex lattice prediction of
subsonic aerodynamics of hypersonic vehicle concepts, J.
of Air., 1977, October.
Pittman, J.L. Application of supersonic linear theory and
hypersonic impact methods to three nonslender hypersonic
airplane concepts at mach numbers from 1.10 to 2.86,
NASA, 1979. TP-1539.
Middleton, W.D.; & Carlson, H.W. Numerical method of
estimating and optimizing supersonic aerodynamic
characteristics of arbitrary planform wings. J. of Air., 1965,
2(4).
Lamb, M.; Sawyer, W.C. & Thomas, L. Experimental and
theoretical
supersonic
lateral-directional
stability
characteristics of a simplified wing-body configuration with
a series of tail arrangements, NASA, August 1981. TP-1878.
Jackson(Jr), C.M.; & Sawyer, W.C. A method for calculating
the aerodynamic loading on wing-body combinations at
small angles of attack in supersonic flow. NASA, 1971. TN
D-6441.
Ehlers, F.E.; Epton, M.A.; Johnson, F.T.; Magnus, A.E. &
Rubbert, P.E. An improved higher order panel method for
linearized supersonic flow. AIAA 16th Aerospace Sciences
Meeting. AIAA. January 1978. pp. 78-15.

169

170

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

10.

11.

12.

13.

14.

15.

16.

17.

18.

19.

20.

21.

Pittman, J.L.; & Riebe, G.D. Experimental and theoretical


aerodynamic characteristics of two hypersonic cruise
aircraft concepts at Mach numbers of 2.96, 3.96, and 4.63.
NASA, TP-1767.
Middleton, W.D. & Lundry, J.L. A computational system for
aerodynamic design and analysis of supersonic aircraft,
part 1-general description and theoretical development,
NASA, 1976. CR-2715.
Sommer, S.C. & Short, B.J. Free-flight measurements of
turbulent boundary layer skin friction, in the presence of
severe aerodynamic heating at Mach numbers from 2.8 to
7.0. NACA, 1955. TN-3391.
Stock, H.W. Ein Verfahren zur Berechnung der
aerodynamischen Beiwerte von Flugkorpen im hohen
Uberschall fur massige Anstell-und Ruderausschlagwinkel
und bei beliebigen Rollagen, Z.Flugwiss, 1976, 24(Heft 4).
Gentry, A.E. & Smyth, D.N. Hypersonic arbitrary-body
aerodynamic computer program (Mark III version), Vol. II.
McDonnel Douglas Corp, April 1968. DAC61552.
Clark, L.E. Hypersonic aerodynamic characteristics of an
all-body research aircraft configuration. NASA, December
1973. TN D-7358.
Penland, J.A. & Pittman, J.L. Aerodynamic characteristics
of a distinct wing-body configuration at Mach 6. NASA,
1985. TP-2467.
Covell, P.F.; Wood, R.M. & Bauer, S.X. Configuration trade
and code validation study on a conical hypersonic vehicle.
AIAA/AHS/ASEE Aircraft Design, Systems and Operations
Conference. September 1988.
Middleton, W.D.; Lundry, J.L. & Coleman, R.G. A system for
aerodynamic design and analysis of supersonic aircraft,
Part-1. NASA, 1980. CR3352.
Shankar,V.; Szema, K. & Bonner, E. Full potential methods
for analysis of complex aerospace configurations. NASA,
1987. CR-3982.
Moore, M.E.; & Williams, J.E. Aerodynamic prediction
rationale for analyses of hypersonic configurations. AIAA 27th
Aerospace Sciences Meeting. AIAA, January 1989. pp.
89-525,
Maughmer,M.; Ozoroski, L.; Straussfogel, D. & Long, L.
Validation of engineering methods for predicting hypersonic

Methods

22.

23.

24.

25.

26.

27.

28.

29.

30.

31.

32.

vehicle control forces and moments, J. of Guid., Co., &


Dyna., 1993, 16(4).
Bonner, E.; Clever, W. & Dunn, K. Aerodynamic preliminary
analysis system II, Part I-Theory. NASA, April 1981. CR165627.
Walker, W.J. & Wolowicz, C.H. Theoretical stability
derivatives for the X-15 research airplane at supersonic
and hypersonic speeds including a comparison with windtunnel results. NASA, August 1960. TM X-287.
Pitts, W.C.; Nielsen, J.N. & Kaattari, G.E. Lift and center of
pressure of wing-body-tail combinations at subsonic,
transonic and supersonic speeds. NACA, 1957. 1307.
Harmon, S.M. & Jeffreys, I. Theoretical lift and damping in
roll of thin wings with arbitrary sweep and taper at
supersonic speeds. Supersonic leading and trailing edges.
NACA, 1950. TN-2114.
Haefeli, R.C.; Mirels, H. & Cummings, J.L. Charts for
estimating downwash behind rectangular, trapezoidal, and
triangular wings at supersonic speeds, 1950. NACA,
TN-2141.
Syvertson, C.A. & Dennis, D.H. A second order shockexpansion method applicable to bodies of revolution near
zero lift, NACA, 1957. 1328.
Allen, H.J. Estimation of the forces and moments acting on
inclined bodies of revolution of high fineness ratio. NACA,
1949. RMA9126.
Perkins, E.W. & Jorgensen, L.H. Comparison of
experimental and theoretical normal force distributions
(including reynolds number effects) on an ogivecylinder
body at mach number 1.98. NACA, 1956. TN-3716.
Dillon, J.L. & Creel(Jr.), T.R. Aerodynamic characteristics at
Mach number 0.2 of a wing-body concept for a hypersonic
research airplane. NASA, 1978. TP-1189.
Dillon, J.L. & Pittman, J.L. Aerodynamic characteristics at
mach numbers from 0.33 to 1.20 of a wing-body design
concept for a hypersonic research airplane. NASA, 1997.
TP-1044.
Dillon, J.L. & Pittman, J.L. Aerodynamic characteristics at
Mach 6 of a wing-body concept for a hypersonic research
airplane. NASA, 1978. TP-1249.

171

PART III
AERODYNAMICS OF RAREFIED GASES

CHAPTER 7
GAS DYNAMICS OF RAREFIED FLOWS
7.1

INTRODUCTION

Tsien1 was the earliest to study the aerodynamics of rarefied


gas flows. The earliest detailed theoretical study of forces and heat
transfer on bodies in free molecule flow and their comparison with
experiments were done by Stalder and his associates2-4. Then
followed review articles by Schaaf and Chambre5 and Schaaf and
Talbot6. These works still serve as basic introduction to mechanics
of rarefied gases.
The characteristics of the flow over a body flying at very
high altitudes depend on the degree of rarefaction of the
atmosphere. As the aerodynamic and heat transfer effects depend
on the type of flow, it is necessary to identify this. Towards this, a
dimensionless parameter called the Knudsen number is
introduced. Knudsen number is defined as the ratio of the
molecular mean free path , of the gas, to the characteristic
dimension of the body d
Kn =

(7.1.1)
d

An alternative is also suggested in which the vehicle length


L and width D are combined as a function of effective angle of attack
to determine d, the characteristic length. For this case, the
relation for Knudsen number becomes
Kn =

(L sin + D cos )

(7.1.2)

176

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Values of as a function of altitude may be obtained from


standard atmosphere tables.
From the kinetic theory of gases, the mean free path is given
by
=

16

RT

(7.1.3)

5 2 p

where is the viscosity of the gas, p, the gas pressure, T, the gas
temperature and R, the gas constant.
From the above equation, one can obtain a relation for the
Knudsen number in terms of the Mach and Reynolds number as
Kn = 1.27

M
Re

(7.1.4)

Based on the Knudsen number, the flow regimes are


arbitrarily defined as
Slip flow
Transition Flow
Free Molecule flow

0.01 Kn 0.1
0.1 Kn 10
Kn 10

Slip flow is still governed by the continuum theory (Navier


Stokes equations), except for the introduction of modified boundary
conditions for the finite velocity, known as the slip velocity, and a
temperature jump at the body surface.
In free molecule flow, molecules reemitted from the body
surface after collision, do not collide with the incoming stream until
very far away from the body. This implies that the free stream
molecular velocity distribution is unaffected by the presence of the
body. Flow phenomena are entirely governed by the moleculesurface interaction, and the kinetic theory of gases is utilised to
analyse the flow.
In transitional flow from free molecular to continuum,
inter-molecule collisions near the body and molecule-surface
collisions are of equal importance, and the theoretical analyses are
quite formidable, and no results of direct aerodynamic interest are
available. Except for the highly computer oriented Direct
Simulation Monte Carlo Technique to analyse the flow in this
regime to get accurate values, approximate, empirical bridging
formulae between free molecular and continuum flows are
invariably made use of.

Gas Dynamics of Rarefied Flows

7.2

FREE MOLECULE FLOW ANALYSIS

Inherent in the definition that the flow is free molecular is


that there are no collisions between the incident molecules and the
surface reflected ones near the body. This allows us to consider the
effects of incident and reflected molecules on the body
independently. Further, we assume that the reflected molecules do
not collide again with any other part of the body, implying that the
body surface is concave everywhere.
In free molecule flow, the molecule-surface interaction
determines the amount of energy and momentum transfer. Towards
this, interaction parameters, one for the energy and the other for
momentum transfer are defined.
7.2.1 Surface Interaction Parameters
The degree of energy equilibrium attained between the
surface element and the molecules before they are reemitted is
expressed in terms of a parameter known as the thermal
accommodation coefficient defined as

dE i dE r
dE i dE w

(7.2.1)

where
dEi =
dEr =
dEw =

energy flux incident on the surface of unit area


in unit time.
energy flux reemitted from the surface of unit
area in unit time.
energy flux of the reemitted molecules if all the
incident
molecules
were
reemitted
in
Maxwellian equilibrium with the surface.

It has been suggested that, if necessary, separate thermal


accommodation coefficients for each degree of freedom, viz.,
translational, rotational and vibrational can also be similarly
defined, as all of them are involved in energy transfer. Experiments
have indicated that vibrational degrees of freedom are not affected
by collision with a surface, while only the translational and
rotational energies are involved.
Similarly, two momentum reflection coefficients are defined,
one for the tangential momentum and the other for normal
momentum transfer.

177

178

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

The tangential momentum reflection coefficient t is defined


by

t =

i r
i w

i r
i

(7.2.2)

where
i

tangential momentum carried to unit area of


surface by incident molecules
tangential momentum carried away from unit
area by reflected molecules
tangential momentum carried away from unit
area by diffusely reflected molecules in thermal
equilibrium with the surface. By definition w = 0.

Normal momentum reflection coefficient n is defined by

n =

pi pr
pi pw

(7.2.3)

where
pi

pr

pw

flux of normal momentum (pressure) incident on


the surface
flux of normal momentum (pressure) carried
away by reflected molecules
flux of normal momentum reemitted if all
incident molecules were reemitted in Maxwellian
equilibrium with the surface.

The case for which = t = n = 1 is called diffuse reflection.


It corresponds to an interaction in which the incident molecules
come to a complete thermodynamic equilibrium with the surface.
The case in = t = n = 0 is called specular reflection. In this type
of reflection, there is no energy and tangential momentum transfer
to the surface. It has been observed that the materials used in space
vehicles are of such a type that the accommodation and reflection
coefficients are close to one.

Gas Dynamics of Rarefied Flows

7.2.2 Forces on an Surface Element in Free Molecule Flow


For a gas in equilibrium, kinetic theory states that the
molecular velocities are specified by the Maxwellian equilibrium
velocity distribution function f. The distribution function f is the
probability that the velocity of a randomly selected molecule will
have velocity components lying in an element of velocity space (dU,
dV, dW ), centered at (U,V,W ). It is given by the expression

f (U,V,W ) dU dV dW =

1
cm

( c m )2

(U 2 + V 2 + W 2 )
dU dV dW

(7.2.4)
Here, cm = most probable molecular speed = 2 R T , where R, is
the gas constant and T, the temperature of the gas. U, V and W are
the velocity components of the random or thermal motion of the
molecules.
To calculate the forces experienced by a body, it is first
necessary to obtain the basic momentum transfer to a differential
elemental area dA and then integrate over the body. The motion of a
body with a velocity - q may be transformed into an equivalent
dynamical problem by considering the body to be at rest and the
gas moving towards it with a velocity q. The distribution of velocities
for the incident molecules relative to an observer moving with the
steady gas velocity q is Maxwellian. Consider an orthogonal
coordinate system such that the axes x and z lie on the plane of the
surface element and the y axis is the inward directed normal. Let
u , v and w be the components of the mass velocity along x, y and
z axes respectively. The molecules approaching the element will
have the thermal velocity components superimposed on the mass
velocity components, viz.,
u = u + U , v = v + V and w = w + W

The number of molecules in the velocity range u to u + du, v


to v + dv and w to w + dw per unit volume is
ni f du dv dw
where, ni is the number density of incident molecules.

(7.2.5)

179

180

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

q
x

Figure 7.1. Coordinate system used for a surface element

The number of molecules with velocity components in the


range u to u + du, v to v + dv and w to w + dw that strike the surface
element of area dA in unit time must lie in a cylinder of height v and
base area dA and is
ni v f du dv dw dA

(7.2.6)

Each molecule that strikes the surface has a momentum


component mu in the x direction where m is the mass of a molecule.
The number of molecules in the velocity range u to u + du, v to v +
dv and w to w + dw which strike the front surface will impart to it a
force in the x direction of magnitude.
m n i uv f du dv dw dA

(7.2 .7)

The total x direction force on the front surface of the element


dA due to incident molecules in all velocity ranges is
+ + +

m ni

u v f du dv dw dA

(7.2.8)

Similarly, the y component of the force is


+ + +

m ni

f du dv dw dA

(7.2.9)

and the z component of the force is

m ni

+ +

v w f du dv dw dA

(7.2.10)

Gas Dynamics of Rarefied Flows

In Eqns. 7.2.8 to 7.2.10, the limits of integration for v are


from 0 to because only the molecules with a velocity component
in the positive y direction will hit the surface.
The vector sum of Eqns. 7.2.8, 7.2.9 and 7.2.10 gives the
total force on the element due to molecules in all the velocity ranges.
We are interested in calculating the components of these forces in a
particular direction. Let ax, ay and az be the direction cosines
between the direction in which the force is required and the x, y and
z axes respectively. Then the component in a particular direction of
the total force on the element due to incident molecules is

dF

= m ni
dA incident

+ + +

(a

u + a y v + a z w )v f du dv dw

(7.2.11)
In a similar manner we can compute the force due to
reflected molecules. It is assumed that the molecules are reemitted
from the surface randomly with a Maxwellian velocity distribution
at a temperature of Tr . The reflected molecules are thought of as a
fictitious gas issuing from the rear side of the surface at a
temperature Tr. The number of molecules reemitting from the
elemental surface of area dA in unit time in the velocity range u to
u + du, v to v + dv and w to w + dw is
n r ( v ) f du dv dw dA

(7.2.12)

where, (-v) is used because only the molecules having a component


of velocity component in the negative y direction will leave the
surface. nr is the number density of the reflecting molecules. The
force on the surface element due to reflected molecules must be
equal to the change in momentum between the initial and final
conditions.
Force =

(initial momentum final momentum) of


reflecting molecules.

The initial momentum is zero as the gas is at rest on the


surface. Therefore, the force in the x direction on the surface due to
reemitted molecules in all the velocity ranges is

m nr

u v f du dv dw dA

(7.2.13)

181

182

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Similarly, the force in the y direction due to reflected


molecules is

m nr

+ 0

v 2 f du dv dw dA

(7.2.14)

while the force in the z direction is


+ 0 +

v w f du dv dw dA
m n r
(7.2.15)


The limits of integration for v are from to 0 as only the
molecules having a velocity component in the negative y direction
will leave the surface.

The sum of the components of the above in the desired


direction having the direction cosines ax, ay and az is
dF

dA

= m nr

reflected

(a

u + a y v + a z w v f du dv dw

(7.2.16)
The sum of Eqns. 7.2.11 and 7.2.16 give the total force on
the surface element in a specified direction due to both the incident
and reemitted molecules and is given by
dF

dA

= m n i

+ + +

(a

+ m nr

u + a y v + a z w )v f du dv dw

(a

u + a y v + a z w ) v f du dv dw

(7.2.17)

The reflecting molecules are having no mass velocity, only


the thermal velocity components. Performing the integration results
in
dF = m n 1 (a u + a v + a w )
i
x
y
z
dA
2
v (1 + erf v ) + 1 e 2 v 2

ay
a

+
(1 + erf v ) + m nr y2
4 2
4 r

(7.2.18)

Gas Dynamics of Rarefied Flows

In the above equation =

r =

, ni and nr are
2 RTi
2 RTr
the number densities of incident and reflected molecules and
x

erf x =

y2

dy

Since the number of molecules impinging on the surface in


unit time must be equal to the number of molecules reflected in
unit time, a relation exists between ni and nr.
The number of molecules in the velocity range u to u + du, v
to v + dv and w to w + dw that strike the unit area of the surface in
unit time is given by
ni v f du dv dw

(7.2.5)

The total number of molecules incident on unit area of


surface in unit time having all possible velocity ranges is given by
+ + +

N i = ni

vf du dv dw

(7.2.19)

Substituting the expression for f, the Maxwellian distribution


function
3

N i = ni
2 R T
i

+ + +

ve

[ (u u )

2 RT i

+ (v v

)2

+ w2

]
du dv dw

(7.2.20)
on integration, it reduces to
v

RT i 2 R T
i
+
e
2

N i = ni

v
2 R Ti

1 + erf

2 R Ti

(7.2.21)
v

The number of molecules in the velocity range U to U + dU, V


to V + dV and W to W + dW that are reemitted from the surface of
unit area, in unit time is

183

184

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

n r ( V ) f dU dV dW

(7.2.22)

These molecules are coming from the surface with no mass


velocity. Hence, their velocity components are due to thermal
motion only. The total number of such molecules in all velocity
ranges is
+

N r = nr

( V )
2 R Tr

3 2

1
2 R Tr

(U 2 +V 2 +W 2 )

dU dV dW

(7.2.23)
The result of the integration is

N r = nr

R Tr

(7.2.24)

Since Ni = Nr , equating the two expressions of Eqns. 7.2.21


and 7.2.24.

n r = n i r v

(1 + erf v ) +

v2

(7.2.25)

Substituting for nr in Eqn. 7.2.18

1
1
= mn i
(a x u + a y v + a z w ) v (1 + erf v ) + e 2 v 2
dA

2
ay
ay
1

2 v 2
+

(1 + erf v ) +
v 1 + erf v + e
2

4
4

dF

(7.2.26)
Let bx, by and bz be the direction cosines between the mass
velocity and the x, y and z axes respectively. Then,
u = b x q , v = b y q and w = b z q

(7.2.27)

Introducing a dimensionless quantity s known as the


speed ratio, which is the ratio of the mass velocity to the most

Gas Dynamics of Rarefied Flows

probable random speed of the molecules (s =

2 RTi

Eqn. 7.2.26 becomes


dF

= mn i

dA

q2
2

{(a x b x

+ a y by + a z b z

1
(b s )2
e y
b y (1 + erf (b y s )) +
s

ay
+
(1 + erf (b y s ))
2s2
+

ay
by
2 r s

(1 + erf (b y s )) +

Since, mni = ri and

1
s

(7.2.28)
e

(b y s )2

Tr

) the

, Eqn. 7.2.28 can be written


Ti

as
dF
dA

= i

q2
2

{(a

x bx

+ a y by + a z b z

ay
1

(b s )2
e y
(1 + erf (b y s ))
b y (1 + erf (b y s )) +
+
2
s

2s
+

Tr b y
1 (b s )2
1 + erf (b y s ) + 2 e y

Ti s
s

ay
2

(7.2.29)

To get the force in the non-dimensional coefficient form we


2
divide the above by q A . The result is
i
ref
2

{(

dC
1
=
dA
A ref
ay
+
2s 2
ay
+
2

a xb x + a yby + a zb z

(1 + erf
Tr
Ti

( ))


b y 1 + erf b y s

(b s ))

( )

bys

b y

(1 + erf
s

(b s )) + s1
y

( )

bys

(7.2.30)

185

186

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

This is an exact expression for the force coefficient under


the assumption of diffuse reflection. All the quantities in the
above equation are known except T r. For this, we have to know
the energy transfer to the surface. If the body is a good thermal
conductor, the surface temperature T w will be the same all over
the surface. If the thermal accommodation coefficient for the
surface material is known then T r can be determined from
Eqn. 7.2.1.
We can remove the restriction of complete diffuse reflection
by introducing the normal and tangential accommodation
coefficients in the force coefficient equation. From Eqn. 7.2.2
r = (1 t ) i

(7.2.31)

and from Eqn. 7.2.3

p r = 1 n

)pi

+n pw

(7.2.32)

The total shear stress on the elemental surface is


= i r = i (1 t

)i

= t i

(7.2.33)

and the total normal stress on the surface is


p = p i + p r = p i + (1 n ) p i + n p w = (2 n ) p i + n p w

(7.2.34)
Substituting the above in the force coefficient Eqn. 7.2.30
we have
dC
dA

1
Aref

2 n a y b y + t a x b x + a z b z

(2 n )a y

1
(b s )2
e y
[1 + erf (by s )]
+
b y (1 + erf (b y s )) +
2s 2
s

n a y
2

Tr b y
1 (b s )2

(
1 + erf (b y s )) + 2 e y
Ti s
s


(7.2.35)

By integrating the results of this relation over the entire


surface of the body, the aerodynamic forces and moments can be
calculated for any arbitrary shaped body. Since a body in free

Gas Dynamics of Rarefied Flows

molecule flow is assumed not to disturb the flow, the complex


vehicle configuration is resolved into a combination of simple subshapes, such as flat panels, cylinders, cones, conical frustums,
spheres and spherical segments. Each sub-shape is subdivided into
an arbitrary number of elemental areas and is integrated over its
area. For each elemental area set up by the integration routine, it is
necessary to determine if the velocity vector intersects the elemental
area directly without first intersecting any portion of the sub-shape
or any of the other composite sub-shapes. If it does, then it is
considered to be in the shadow region and this has to be treated
appropriately.
7.3

AERODYNAMIC FORCES FOR TYPICAL BODIES

Aerodynamic force coefficients in free molecule flow for


simple body shapes can be determined by the application of the
results of the previous section. In the analysis it is assumed that
the body temperature Tw , the tangential and normal momentum
accommodation coefficients and the temperature Tr of the molecules
reflected from the surface are all constant over the entire surface.
7.3.1 Flat Plate
CASE 1: ONLY THE FRONT SIDE EXPOSED TO THE FLOW
The flat plate is at an angle of attack a to the incoming
velocity q. The elemental surface coordinates x, y and z are as
shown. It is to be remembered that while deriving the basic relation
for the force on an elemental surface, positive y axis was specified
as inward normal to the surface. X and Y are the directions in which
the force coefficients are sought, viz., the tangential and normal
force coefficients.

Y
y
j
i

z OUT OF PAGE

z OUT OF PAGE

Figure 7.2. The flat plate exposed on the front side only

187

188

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

The force coefficient from Eqn. 7.2.35 is


dC
dA

1
Aref

2 n

a b + a b + a b
y y
t
x x
z z

b y 1 + erf b y s +

2 a
n

y
+
2
2s
n a y
2

1 + erf

1
s

b y s

b s +
y

b y s
Tr b y
1

1 + erf b y s + 2 e

Ti S
s

(7.3.1)
In the above ax, ay and az are the direction cosines between
the local x, y and z axes and the desired force direction, bx, by and bz
are the direction cosines between the local x, y and z axes and the
mass velocity vector.
The unit vector in the direction of the mass velocity is
q = i cos + j sin

For the front surface


x = i ; y = j and z = k

The direction cosines between the mass velocity and the x, y


and z axes are
bx = q . x = cos ; by = q . y = sin and bz = q . z = 0
Normal Force
For the force in the normal direction, the direction cosines are
ax = x . j = 0; ay = y . j = 1 and az = z . j = 0
Substituting the values of ax , ay , az and bx , by , bz in
Eqn. 7.3.1 and basing the reference area as the area of the front

Gas Dynamics of Rarefied Flows

surface of the flat plate, the expression for the normal force
coefficient is
dC N front = (2 n

sin e ( s sin )
1
sin 2 +
(1 + erf (s sin )) +

2s 2
s

Tr

2s 2

Ti

[e

( s sin )2

s sin (1 + erf (s sin ))

(7.3.2)
Axial Force
The axial force is in the X direction. The direction cosines
between the x, y and z axes and the X axis are given by
ax = 1, ay = 0 and az = 0
as before
bx = cos , by = sin and bz = 0.
Substituting the above in Eqn. 7.3.1, we obtain for the axial
force coefficient as

dC A front

= t sin cos

CASE 2:

e (s sin )

+ (1 + erf s sin )
s sin

(7.3.3)

REAR SIDE OF THE FLAT PLATE ONLY


EXPOSED TO THE FLOW
Y
Z IN TO PAGE

j
x
i
X

Z OUT OF PAGE

q
y

Figure 7.3. Flat plate exposed on the rear side only

The elemental coordinate system for this case is as shown,


with the y axis pointing downwards (inward normal to the rear
surface). As before, X and Y are the directions of axial and normal
forces sought respectively.

189

190

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

x = i ; y = j and z = k

Normal Force
The direction cosines in this case are
ax = 0, ay = -1 and az = 0
and
bx = cos , by = - sin and bz = 0
Substituting the values in Eqn. 7.3.1, we obtain the normal
force coefficient as
dC N rear = (2 n

sin e (s sin )2
1

sin 2 +

2 s2
s

Tr

2s 2

Ti

[e

( s sin )2

(1 erf (s sin ))

s sin (1 erf (s sin ))

(7.3.4)
Axial Force
The corresponding direction cosines in this case are
ax = 1, ay = 0 and az = 0
and
bx = cos , by = - sin and bz = 0.
Substituting in the Eqn. 7.3.1 we get

dC Arear

CASE 3.

e (s sin )

= t sin cos
(1 erf (s sin ))
s sin

(7.3.5)

BOTH SURFACES OF THE PLATE EXPOSED TO


THE FLOW

dC N = dC N front + dC N rear

Adding the expressions of Eqns. 7.3.2 and 7.3.4 we get the


normal force coefficient as

Gas Dynamics of Rarefied Flows

dC N

= (2 n )

1
2 sin e ( s sin )

2 sin 2 + 2 erf (s sin ) +


s
s

+ n sin

Tr
Ti

(7.3.6)
From Eqns. 7.3.3 and 7.3.5, the axial force coefficient is given
by

dC A

e (s sin )
= 2 t cos sin erf (s sin ) +

(7.3.7)

If one is interested in the lift and drag coefficients, they can


be determined by the relation
CL = CN cos - CA sin
CD = CN sin + CA cos
7.3.2 INFINITE RIGHT CIRCULAR CYLINDER AT AN ANGLE
OF ATTACK,
The cylinder being axially symmetric, we can orient the body
axis system in such a way that the velocity vector is in the XY
plane.
The elemental area chosen for integration is
dA = r d dL
where, dL is the elemental length along the cylinder symmetry axis.
A reference body coordinate system X, Y and Z is chosen such that
the axes X and Y represent the directions of desired tangential and
normal forces respectively and Z axis normal to them as shown. Let
x, y and z be the local elemental surface coordinates with y axis
radial inwards and x and z axes normal to each other and tangential
to the surface element.
The unit vector in the direction of the mass velocity is
q = i cos + j sin

191

192

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows


Y
j
y

k
Z

Figure 7.3. Infinite right circular cylinder at an angle of attack,

The unit vectors in the direction of the elemental area axes


system are
x =i
y = j cos k sin
z = j sin + k cos

The direction cosines bx, by, bz are


bx = q . x = cos
by = q . y = sin cos
bz = q . z = sin sin
Normal Force
For the normal force, the direction coefficients are
ax = x . j = 0
ay = y . j = cos
az = z . j = sin
The integration limits for are from 0 to 2.

C N cyl . =

1
Aref

2 L

(dC ) r d dL

Substituting the direction cosine values in the expression


for dC from Eqn. 7.3.1 and integrating over the surface area of the
cylinder, the normal force coefficient for the cylinder, based on the
frontal projected area 2rL as the reference area Aref is

Gas Dynamics of Rarefied Flows

CN =

n sin
4s
e

Tr
Ti

s 2 sin 2
2

+ sin 2 (4 + t 2 n

f (s sin )

(7.3.8)

where

1
1

+ s sin I o
f (s sin ) =

2
3
s
sin

1
s sin

+
+
I1
3
6 s sin

and

s 2 sin 2

s 2 sin 2

(7.3.9)

s 2 sin 2

Io
= Bessel function of first kind and zero order
2

s 2 sin 2

I1
= Bessel function of first kind and first order
2

Axial Force
The direction cosines for the axial direction force are
ax = x . i

=1

ay = y . i

=0

az = z . i

=0

Substituting the direction cosine values in the expression


for dC in Eqn. 7.3.1, and integrating over the cylindrical surface
gives the result

C Acyl. =

(s sin )2
2

(s sin )2
cos
2
1 + (s sin ) I o
2

(s sin )2
2
+ (s sin ) I 1

(7.3.10)
where, Io and I1 are Bessel functions mentioned earlier

193

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

7.3.3 SPHERE
An expression for the drag coefficient of an element of area
dA when both of its sides are exposed to the flow can be obtained
from the normal and axial force relations for a flat plate discussed
earlier, by the relation
CD = CN cos + CA sin
Multiplying Eqn. 7.3.6 by cos and Eqn. 7.3.7 by sin ,
and adding the two expressions we get

dC D =

1
Aref
+

2 e (s sin )2

sin 2
s

(2 n

[ (2

Tr
Ti

) sin 2 +

) sin 2 + t cos 2 ]

+ [2 sin erf (s sin ) ]

1
+ t cos 2 dA
2
2s

(7.3.11)
Element of area for sphere dA = 2R2 cos d
Choose an elemental ring surface of area dA = 2 R 2cos
on the front and rear side of the sphere as shown in Fig. 7.3.4. Each
d

R cos

194

Figure 7.3.4. Figure showing drag coefficient of an element of area dA


exposed on both sides to the flow.

Gas Dynamics of Rarefied Flows

point on this ring element will be at the same angle of attack and
the expression for dCD can be applied. In general, the temperature
Tw varies from one surface element to another. In the special case
where the body is a perfect thermal conductor, Tw is constant over
the entire surface, the above expression can be integrated over the
surface to get the drag coefficient of the sphere based on the frontal
projected area.

C D sph ere

1
=
A ref

dC

dA

(7.3.12)

The result of the integration is

C D sphere

2 n + t
=
s3

2 n
3 s

4 s 4 + 4 s 2 1
e s

erf s +

4s

2 1
s +
2

Tr
Ti
(7.3.13)

7.3.4 CONE FRUSTRUM


Closed form solutions for the force coefficients are not
available for other body geometries and one has to go in for
numerical integration. The analysis for a cone at an angle of attack
has been outlined in detail5,7. The final results for the normal and
axial force coefficients taken from Sentman7 is reproduced below. It
is assumed that the accommodation coefficients n and t are equal
to one.
L2

L1

j
y

x
Figure 7.3.5. Figure depicting normal and axial force coefficients for
a cone frustrum.

195

196

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

We consider only the conical surface omitting the front and


rear ends of the cone.
From the figure
r = L tan and dr = dL tan
The element of area considered is
dA = r d ds
(ds)2 = (dL)2 + (dr)2 = (dL)2 (1 + tan2 ) = (dL)2 sec2
ds = sec dL
dA = L

tan
cos

dL d

The unit vectors in the direction of the local axis system are
x = j sin k cos
y = i sin + j cos cos k cos sin
z = i cos j sin cos + k sin sin

The body is oriented in such a way that the velocity vector is


in the X-Y plane.
The unit vector in the direction of mass velocity is
q = i cos + j sin

The direction cosines bx, by and bz are


b x = x . q = sin sin
b y = y . q = cos sin + sin cos cos
b z = z . q = cos cos sin sin cos

Normal Force
For the normal force the direction cosines are
a x = x . j = sin
a y = y . j = cos cos
a z = z . j = sin cos

Gas Dynamics of Rarefied Flows

Substitution of the above direction cosine values in the Eqn.


7.3.1 and integrating the resulting expression for from 0 to 2 we
get for the normal force coefficient
L 22 L12 tan

C N cone =

cos

2 Aref

{2 sin cos sin + 2 sin cos sin

erf (b y s ) d + cos 2 sin 2 + 2 cos erf (b y s ) d


s o

2 sin
s

(b y s )2

d +

3
Tr 2
sin cos 2 +
Ti 2s

cos
s

(cos sin cos erf (b y s ) d


o

+ sin cos cos erf (b y s ) d ) +

cos
s

cos e

(b y s )2

(7.3.14)
Axial Force
For this case the direction cosines are
ax = x . i = 0
a y = y . i = sin

a z = z . i = cos

Substituting the above in Eqn. 7.3.1 and integrating for


between the limits 0 to 2 we get for the axial force coefficient

C Acone

L 22 L12 tan
1

2
sin 2 cos + 2
2 Aref cos
s

+ sin 2 cos 2 + 2 erf (b y s ) d


s o

+ 2 sin cos cos cos erf (b y s ) d +

2 cos
s
Contd ...

197

198

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

(b y s )2

d +

sin

Tr 3 2
cos sin 2 +

Ti s

(cos sin

erf (b s ) d + sin cos cos erf (b s )d )


y

sin
s

(b y s )2

(7.3.15)

7.3.5 SPHERICAL SEGMENT


Most of the space vehicles have spherical caps in front of the
body. It is therefore useful to have relations about the normal and
axial pressure coefficients for a spherical segment. The details of
the analysis are given by Sentman7 and the final results from it are
reproduced below. The back side of the spherical segment is not
considered. The element area of the sphere chosen for integration is

dA = r

sin d d

The integration limits are


0 to 1 for
and 0 to 2 for .
Substituting the relevant direction cosines in the expression
for the force on an element Eqn. 7.3.1 one obtains the following
expressions for the normal and axial force coefficients.
The normal force coefficient is given by

Gas Dynamics of Rarefied Flows

Y
j

Figure 7.3.6. Normal and axial pressure coefficients for a spherical


segment.

CN =

sin
cos
sin
2
sin
cos
sin cos erf by s dd

1
Aref
o o

r2

+ 2 sin 2 +

s2
+

2 sin
s

sin e

sin

cos erf b y s d d

o o

by s 2

d d

o o

cos 3 1
2
Tr 3 2

cos 1 +
sin

3
3
Ti 2 s


cos sin 2 cos erf b y s d d
cos
s
o o

+ sin

1
s2

sin 3 cos 2 erf b y s d d

o o

(b y s )2

sin 2 cos e
d d

o o

where, by = cos cos + sin sin cos

(7.3.16)

199

200

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

The axial force coefficient is given by

C A sph . sg

r2

2
2
sin 2 1
=
cos sin 1 +
2
Aref
2s
1

+ 2 cos 2 + 2
s

sin cos erf (b s )d d


y

o o

+ 2 sin cos

sin

cos erf (b y s )d d

o o

+
+

2 cos
s

sin e

d d

o o

Tr 3 2
cos 1 cos 3 1

Ti 3 s

(cos sin cos

+ sin

sin
o

1
s2

erf (b y s ) d d

(7.3.17)

o o

(b y s )2

cos cos erf (b y s )d d )

o o

sin cos e

(b y s )2


d d

where, by = cos cos + sin sin cos as before.


7.4

AERODYNAMIC FORCES IN SLIP & TRANSITIONAL


FLOWS

Theoretical methods based on Thirteen Moment and


Burnett equations to approximate the integro-differential equation
of Boltzman have been utilised to analyse the flow problems in the
near free molecule or transitional flow conditions. Only a few special
cases such as shock structure, couette flow, etc., have been studied
and these are of no practical use in the determination of the
aerodynamic forces or heat transfer of bodies in this flow regime.
Computer oriented Direct Simulation Monte Carlo Technique is the
only known method to give reliable quantitative aerodynamic force
and heat transfer parameters all the way from free molecule to

Gas Dynamics of Rarefied Flows

continuum. The involved computer programming, with separate


analysis for each flow condition does not allow this method to be of
use in the preliminary analysis stage.
Simple empirical formulae have been suggested for quick
determination of the flow parameters in the transitional flow
regime. It has been experimentally observed that the aerodynamic
force parameters vary monotonically from free molecule to
continuum flows. Utilizing this aspect, several authors have
suggested so-called bridging formulae.
Some of these are,

C FTRAN = C FCONT

1
1
2

+ C F FM C FCONT sin

+
log 10 Kn
3

(7.4.1)
where, CF is the force coefficient and the subscripts TRAN, CONT
and FM refer to transition, continuum and free molecule flow
regimes respectively. The continuum values are usually based on
the Newtonian analysis. The above equation and its correlation with
the experimental data are discussed by Lott8.
Another similar empirical bridging formula suggested by
Blanchard9 is
C FTRAN = C FCONT + C F FM C FCONT sin 2 w

(7.4.2)

where
w = ( 3 + log 10 Kn ) / 8

Based on considerations of fluid dynamic simulation


appropriate to hypersonic viscous flows over blunt nosed bodies,
Potter10, 11 has suggested a method for estimating the lift and drag
coefficients in the transitional flow regime. According to him, a
simulation parameter suitable for scaling viscous, hypersonic flow
effects was devised. The characteristic length in this parameter was
taken to be a shape factor modified wetted length in the streamwise direction. To quote Potter, this approach was suggested by the
empirical data showing that projected area and wetted area override
detailed body shape variations in determining overall forces in
hypersonic, transitional flows.

201

202

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

The suggested method for correlating the lift and drag


coefficients in hypersonic rarefied transitional flows are as follows:
A normalized form of drag coefficient is defined as
C D C Di

CD =

(7.4.3)

C D fm C D i

For a variety of practical vehicle shapes, the above is


correlated with a simulation parameter defined as

Pn D

H
s * *

(7.4.4)

where
U = free stream velocity
= kinematic viscosity
1

s * =s (P F A /W A )2

= 0.63

H
H
s

H
( 0.2 H o + 0.5 H w )

= streamwise wetted length

PFA = projected frontal area


WA = wetted area
2nD = drag parameter

CD = drag coefficient

C D i = inviscid drag coefficient


Similarly for the lift case
_

CL =

C L C Li
C L fm C L i

2nL = 2nD (PPA /PFA)1/ 4, the lift parameter

where, PPA is the planform projected area.

(7.4.5)
(7.4.6)

Gas Dynamics of Rarefied Flows

The available experimental data of lift and drag in


transitional flow was normalised as above and plotted against 2nL
and 2nD . From the graph it was found that one analytical curve was
able to adequately fit the lift and drag data. A simple form of such a
curve is
_

CL

CD

2. 6
=
2 . 6 + Pn1L.6

2. 6
=
2 . 6 + P 1. 6
nD

0.5

0.5

(7.4.7)

(7.4.8)

In view of the apparently good agreement with full scale


flight data, Potter10, feels that Eqns. 7.4.7 and 7.4.8 represent
useful tools for preliminary design studies.
7.5

ENERGY TRANSFER IN FREE MOLECULE FLOW

The molecular energy transfer with the surface element


involved by the incident and reflected molecules can be determined
in a manner very similar to that of momentum transfer studied
earlier.
Incident energy on the front side of an element
The translational energy incident on the front side of an
element of are dA due to incident molecules in unit time is

(dE )
i tr

1
2

+ + +

n i u 2 + v 2 + w 2

v f u ,v , w

du dv dw dA
(7.5.1)
The result of the integration is

dE = i R T i

i tr

f
2

3 2

5
+ s2 +

2
s + 2

2
e s (s sin )

s sin 1 + erf s sin dA

)]

(7.5.2)

203

204

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

If the gas is monotomic, then the total incident energy is


given by the above expression. However, if the gas is composed of
diatomic molecules, then each molecule carries an additional
amount of energy called the internal energy. By the principle of
equipartition of energy, the amount of energy carried by each
molecule is () j mRTi, where j is the number of degrees of freedom.
j is related to the ratio of specific heats by the relation.
j = (5 3 )/( 1)

(7.5.3)

In contrast to the translational and rotational degrees of


freedom, the process of vibrational energy exchange is inefficient as
thousands of collisions are required for effective energy transfer.
For this reason, at normal temperatures the translational and
rotational degrees of freedom are considered active, and vibration
as an inert degree of freedom. At normal temperatures, the value of
j is approximately 2 for air.
The internal energy of molecules striking the front surface
in unit time is
j
m R T i N i f dA
( dE int ) f =
(7.5.4)
2
where, Nifront is the number of molecules striking the front surface in
unit time and is given by the Eqn. 7.2.21, viz.,

{e
2

RT i

N i f = ni

( s sin )2

(s sin )

( 1 + erf s sin ) dA
(7.5.5)

Subsituting Eqn. 7.5.5 in Eqn. 7.5.4 we have

( dE int ) f =

j
2

{e

m (R T i

( s sin )2

ni

)3 2

2
(s sin )

( 1 + erf s sin ) dA

(7.5.6)

The total incident energy is (dEi)f = (dEtr)f + (dEint)f

(dE )
i

(RT )

3 2

= i

( s sin )2
e

s sin

s2 + 2 + j

) (1 + erf s sin ) s

5
2

j
dA
2

(7.5.7)

Gas Dynamics of Rarefied Flows

Similarly, for the rear surface

(dE i )r = i

(RT i )3 2
2

(s sin )

( s sin )2

j
2
s + 2 + 2

( 1 erf s sin ) s 2 + + dA
2 2

(7.5.8)

Energy carried by the reflected molecules


The energy dEw carried by the reflected molecules that are in
Maxwellian equilibrium with the surface can be calculated in a
similar manner.
+

dE
= 1mn

w tr
wf

f
2

+V

+W

V f U ,V ,W

dU dV dW dA
(7.5.9)
On integration
dE
= 2 w
w tr

(RT )
w

3 2

dA

(7.5.10)

From Eqn. 7.2.25

wf = i

T i (s sin )2
+
e
Tw

)]

s sin 1 + erf s sin

(7.5.11)

Substituting for w in Eqn. 7.5.10 we get


f

3
dE = 2 i R

r tr

f
2

Tw f

Ti

)]

(s sin )2

+ s sin 1 + erf s sin dA


e

(7.5.12)

The internal energy of reemitted molecules is

205

206

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

(dE )
int

(7.5.13)

m R T w f N w f dA

Since N w f = N w i , the above reduces to

j i R3

(dE int ) f =

{e

Tw f

2 2
( s sin

)2

Ti

s sin [1 + erf (s sin )] dA

(7.5.14)
The total energy of reflected molecules in equilibrium with
the surface is

(dE w )f
tr

+ dE w int

)f

= (dE w ) f

3
j i R

(dE w ) f = 2 +
2

Tw f

(7.5.15)

Ti

{e

( s sin )2

(7.5.16)

+ s sin [ 1 + erf (s sin )] dA


Similarly, for the rear side of the surface

3
j i R

(dE w )r = 2 +
2

{e

( s sin ) 2

Tw r

Ti

s sin [ 1 erf (s sin )] dA


(7.5.17)

For specular reflection, the accommodation coefficient = 0,


and there is no energy transfer to the body. If 0 and if there is no
internal and radiation heat transfer, then the elemental surface will
reach a temperature known as the Equilibrium Temperature. This
value can be determined by equating the incident energy to the
reflected energy.

Gas Dynamics of Rarefied Flows

7.5.1 Equlibrium Temperatures for Simple Shapes


Flat Plate: Front surface only exposed to the flow
The energy relations derived above for an elemental surface
refers to the case of a flat plate at angle of to the free stream. The
following expression gives the ratio of the equilibrium temperature
to the free stream temperature when only the front surface is
exposed to the flow.

(T )
wequ

Ti

1
j

2 +
2

j (s sin )2 2 5 j

2
+ s + + s sin [1 + erf (s sin )]
s + 2 + e
2
2
2

( s sin )2
e

+ s sin [1 + erf (s sin )]

(7.5.18)

Similarly for the rear surface

(T )

wequ r

Ti

1
j

2 +
2

j (s sin )2 2 5 j
2

s + + s sin [1 erf (s sin )]


s + 2 + 2 e
2 2

(s sin )2
s sin [1 erf (s sin )]
e

(7.5.19)

When both the front and rear surfaces are in perfect thermal
contact
Twequ
1
=
j
Ti

2 +
2

j (s sin )2 2 5 j
2

+ s + + s sin erf (s sin )


s + 2 + 2 e
2 2

2
(s sin )
+ s sin erf (s sin )
e

(7.5.20)

207

208

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Sphere
If the temperature Tw is constant over the entire surface of
the sphere then the expressions for incident and reflected energies
on an element of the sphere can be integrated to obtain the
equilibrium temperature. The elemental area chosen is identical to
the momentum calculation case. The result of the integration is

Tweq .
Ti

1
=

j
2+
2

j 1 3 j
2 5 j s2 3

s + + e
+ s + s 3 + + +
2 2
2 s 4 4

2
1

erf (s )
e s + s +
s
2

erf (s )

(7.5.21)
For a monatomic gas j = 0 and j =2 for a diatomic gas.
Cylinder
For the case of a cylinder having a constant surface
temperature, the energy transport equations for an elemental area
can be applied and integrated over the cylinder surface to get the
equilibrium temperature. The elemental area chosen for integration
is similar to the case of momentum transfer. The result is

Tweq
Ti

s2 4 2 7 j
s2
j
5s2 j s2
+
I o s + s + + 2 + + I 1 s 4 +

2
2
2
2
2 2
1 2

=

j
s2 2
s2 2

2+
I o s +1 + I 1 2 s

2
2
2

(7.5.22)

where, Io and I1 are modified Bessel functions of first kind, zero and
first order respectively.
7.5.2 Heat Transfer for Typical Bodies in Free Molecule Flow
The two non-dimensional parameters that are normally
used in heat transfer calculations are the thermal recovery factor
and the Stanton number. They are defined by
r =

Tw eq . Ti
Ttot . Tw

(7.5.23)

Gas Dynamics of Rarefied Flows

St =

A i V i c p Tw eq . Tw

(7.5.24)

where, Q is the total heat transfer, Ttot. is the adiabatic stagnation


temperature, Ti is the free stream temperature, Tw is the body
surface temperature, pi is the free stream density, Vi is the free
stream velocity and cp is the specific heat at constant temperature
and A is the total heat transfer surface area.
dQ = (incident energy reflected energy ) = dEi - dEr
The thermal accommodation coefficient from Eqn. 7.2.1 is

dE i dE r
dE i dE w

From the above


dQ = dEi dEr = (dEi dEw)
Eqns. 7.5.7 and 7.5.16 give the expressions for dEi and
dEw for a front-surface element.
When dQ is zero

Tw = T eq.

Eqn. 7.5.18 gives the expression for the equilibrium


temperature of a surface element.
The Stanton number can be obtained by integrating the
dQ
expression
over the total surface of the body of interest.
Teq . T w
Similarly, for the thermal recovery factor. The results thus obtained
for various simple shaped bodies are given by Schaff & Talbot6 and
are presented below.
Front Face of a Flat Plate at an Angle of Attack

r =

( + 1)s

2
2s + 1

1+

( s sin )[1 + erf (s sin )]e (s sin )


(7.5.25)
1

209

210

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

{e

( +1)

St =

4 s

( s sin 2 )

(s sin ) [1 + erf (s sin )]

(7.5.26)

Rear Side of the Flat Plate at an Angle of Attack


For the case of the rear side of the flat plate exposed to the
flow, the expressions are obtained by replacing by in the above
equations, viz.,

r =

( + 1)s 2

1
2
2s + 1
2

1 ( s sin ) [1 erf (s sin )]e (s sin )

(7.5.27)
St =

( + 1 )
4 s

{e (

s sin )2

(s sin ) [1 erf (s sin )]

(7.5.28)
Front and Rear Surfaces of the Flat Plate
It is assumed that the front and rear surfaces are in perfect
thermal contact and the plate temperature Tw is the same
throughout.

r =

( + 1)s 2

2
2s + 1

1+

( s sin ) erf (s sin ) e (s sin )


1

(7.5.29)
St =

{e

( + 1)
4 s

( s sin )2

(s sin ) erf (s sin )

(7.5.30)

Infinite Circular Cylinder at an Angle of Attack

r =

( + 1)s 2

2
2
2
2

+ 2 1 + 2s 2
s
I

+
+

2
1

2
2

I
I
+

1
1
o

2
2

)]

[ (

)] I 1 2

(7.5.31)

Gas Dynamics of Rarefied Flows

St =

( + 1 )

4s

2
2

2
2
2
2

+ I1
1+ I o

2
2

(7.5.32)

2
2
and I 1
are modified Bessel
where, = s sin and I o

2
2
functions of the first kind, zero and first order respectively. The
area considered being the total curved surface area L d where L is
the length of the cylinder.
Sphere

2
2s + 1

r =

+1
s2

1
erf ( s )

2s

1
1

(
)
erf
s
1 + ierfc (s ) +
2

2s

) 1 + 1s ierfc (s ) + 2s

(7.5.33)
St =

( +1 )
8s

2
s + s ierfc ( s

)+

1
2

erf

( s )

(7.5.34)
where, ierfc = {1 erf (s)} dA is the complimentary error function
integrated over the total heat transfer area, d 2.
Cone at an angle of attack
Only the conical surface considered neglecting the base.

Q = L 2 i R Ti
+

R Ti tan
2 cos

+ 1 Tw
2
s +

1 2 ( 1 ) Ti

(s sin ) [1 + erf (s sin )]

2
e ( s sin )
2

( s sin )2
e

(7.5.35)

where, L is the height of the cone, the semi-vertex angle of the


cone and
sin = cos cos sin + cos sin

211

212

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Integrating the above and substituting the value in the


Eqn. 7.5.24, the Stanton number can be obtained.
When there is no heat transfer, by equating Q = 0, the
equilibrium temperature is obtained, as Tw will be equal to T w ( eq .)

Tw ( eq . )
Ti

=1 + r

( 1)
(7.5.36)

s2

From the above relation the recovery factor can be


determined.
7.5.3 Heat Transfer In Slip & Transitional Flow Regimes
Due to the complexity of interactions between the viscosity,
rarefaction and compressibility, it has not been possible to
theoretically analyze the heat transfer parameters in the slip and
transitional flow regimes, although some approximate theories have
been advanced for limited cases particularly in the slip flow regime.
The Navier-Stokes equations with modified boundary conditions for
slip and temperature jump are the best theoretical approach in the
slip flow regime. Only the computer oriented Direct Simulation
Monte Carlo technique is able to give the aerodynamic and heat
transfer parameters all the way from free molecular to continuum.
For approximate preliminary analysis, recourse is made to empirical
relations.
One such relation is listed below.
The near free molecular heat flux (qnfm ) to a surface is given
in terms of the free molecular heat flux by the following relationship
described by Caruso and Naegeli12

q nfm
q fm

Tw
= 1 + 2

0.1414 s
2
s

Tw

12

Kn

(7.5.37)

where the subscript nfm refers near free molecule or transitional


flow conditions.
For the recovery factor in the slip and transitional flow
regime the suggested formula is

Gas Dynamics of Rarefied Flows

r =

rc + Kn r fm
1 + Kn r fm

(7.5.38)

where, rc is the recovery factor in continuum flow and rfm the


recovery factor in free molecule flow.
REFERENCES
1.
2.
3.

4.

5.

6.

7.

8.

9.

10.

11.

Tsien, H.S. Superaerodynamics mechanics of rarefied gases.


J. Aero. Sci. 1946, 13(12).
Stalder, J.R. & Jukoff, D. Heat transfer to bodies travelling at
high speed in the upper atmosphere. NACA, 1949. Rep. 944.
Stalder, J.R.; Goodwin, G. & Creager, M.O. A comparison of
theory and experiment for high speed free molecule flow.
NACA, 1950. TN 2244.
Stalder, J.R. & Zurik, V.J. Theoretical aerodynamic
characteristics of bodies in free-molecule flow field. NACA,
July 1951.TN- 2423.
Schaaf, S.A. & Chambre, P.L. Flow of rarefied gases, Section
H, In Fundamentals of Gas Dynamics, Vol. III. High Speed
Aerodynamics and Jet Propulsion, Princeton University
Press, 1958.
Schaff, S.A. & Talbot, L. Mechanics of rarefied gases In
Handbook of Supersonic Aerodynamics, Section 16.
NAVORD, 1959. Report 14885.
Sentman, L.H. Free molecule flow theory and its application
to the determination of aerodynamic forces. Lockheed Missile
& Space Company Report, October 1961. LMSC - 448514.
Lott, R.A. Aerodynamic characteristics for the saturn SA-6
vehicle and the SA-5 after orbital breakup. Lockheed Missiles
and Space Company, Huntsville, Ala. Feb. 1964. LMSC/
HREC TM 54/01-44.
Blanchard, R.C., Rarefied flow lift-to-drag measurements of
the shuttle orbiter, Presented at the 15th ICAS Conference,
London, 1986. Paper No. ICAS 86-2.10.2.
Potter, J.L. Transitional, hypervelocity aerodynamic
simulation and scaling. AIAA 20th Thermophysics Conference,
Williamsburg, VA, 1985.
Potter, J.L. Procedure for estimating aerodynamics of three
dimensional bodies in transition flow. In Rarefied Gas
Dynamics. Edited by Muntz, E.P., Weaver, D.P. & Campbell,
D.H. Progress in Astronautics and Aeronautics. Vol. 118.

213

214

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

12.

Caruso, P.S. Jr. & Naegeli, C.R. Theoretical and empirical low
perigee aerodynamic heating during orbital flight of an
atmosphere explorer. Proc. of the 1976 Heat Transfer and
Fluid Mechanics Institute, Stanford University Press.

APPENDIX

THE MARK N SUPERSONIC HYPERSONIC ARBITRARY BODY


The Mark IV Supersonic-Hypersonic Arbitrary Body
Program known as SHABP or HABP for short, is a digital computer
program package that is capable of calculating the supersonic
hypersonic aerodynamic characteristics of complex arbitrary
3-dimensional shapes. This program was developed by A.E.
Gentry and his group at Douglas Aircraft Corporation. In
literature it is sometimes referred to as Gentry's Program.
Complete documentation of the work is presented in three
volumes as Wright-Patterson Air Force Base Flight Dynamics
Laboratory, Technical Reports, (AFFDL-TR-73-159, Vols. 1,2 and
3 , November 1973). The outstanding features of this program are
to quote from its documentation page:
"Itsflexibility in covering a wide variety of problems and the
multitude of program options available. The program is a
combination of techniques and capabilities necessa ry in performing
a complete aerodynamic analysis of supersonic and hypersonic
shapes. These include: vehicle geometry preparation: computer
graphics to check out the geometry; analysis techniques for defining
vehicle component flow field effects; surface streamline
computations; the shielding of one part of the vehicle by other;
calculation of surface pressures using a great variety of pressure
calculation methods including embedded flow field effects; and
computation of skin friction forces and wall temperature. Although
the program primarily uses local-slopepressure calculation methods
that are most accurate at hypersonic speeds, its capabilities have
been extended down into supersonic speed range by the use of
embedded flow field concepts. This pennits the first order effects of
component interference to be accounted for."
The functional organization of th; program is given in
Fig. A. 1. The program is written in Fortran.

2 16

Aerodynamic Predictive Methods and their Validationin Hypersonic Flows

Based on the program formulation and listing given in the


above mentioned AFFDL reports, the aerodynamics group at
DRDL, Hyderabad, has replicated the program and made it
operational so that it could be used for its in-house work. To test
that the replicated program is a copy of the original SHABP
program of Gentry et al., and is capable of giving the desired
results, a test case was chosen. Wind tunnel measurements of
normal force and pitching moment coefficients of body alone, wing
alone and wing body combination at Mach No. 6 and their
comparisons with theoretical predictions have been reported in
NASA TP 2467. Figure A-2 gives details of the model tested and
Figures A-3 to A-8, the measured test results and the comparison
of the same with the theories. In these figures three theoretical
curves are drawn and compared with the experimental data. The
theoretical curves are:
(a) Theoretical curve from the earlier version of SHABP (known
a s the Mark I11 version) a s given in the NASA TP 2467.
(b) Theoretical curve a s calculated from the DRDL program
which is a duplicate of Mark IV version.
(c) Theoretical curve from the DRDL developed program based
on the tangent cone method for the body and tangent wedge
for the wing.
Close agreement with the Mark I11 version results and the
DRDL developed SHABP based on the original Mark IV version of
SHABP shows that the DRDL replicated programme is a true
duplicate of the original version and could be used with
confidence for preliminary aerodynamic analysis. Figures A1-A8
are given on the following pages.

EXECUTIVE
PROGRAM
SERVICE ROUTINES,
ERROR HANDLING, ETC.

READ SURFACE
PROPERTY DATA
CALCULATE
STREAMLINE
TRAJECTORIES

CALC
FRIC

INTEG
FRICT

SURFACE PRO
PERTY DATA SAVE

SKIN FR
FORCE
INTEGRA

FORCE DAT

Figure A-l. Functional organisation of Mark IV progr

2 18 Aerodynamic Predictive Methods a n d their Validation in Hypersonic Flows

Figure A-2. Wing body confirmation as given in NASA TP-2467

Appendix 21 9

7
SHABP (DUD\)

-0.04 - - -

I
-0.08

-15.0

-1d.O

-5.0

0.0

I
I

5.01

ANGLE OF ATTACK (deg)

Figure A-3.Normal force coemcient (body alone) at Mach 6.0

10.

220 Aerodynamic Predidive Methods and their Validationin Hypersonic Hows

ANGLE OF ATTACK (deg)

Figure A-4. Pitching moment coefficient (body alone) at Mach 6.0

Appendix 221

0.20

----

I
I
0I

---h&kFor---Te-I
I
I
I
I
I

-I-

- - -f-

1
I
I

EXPT (TP246Y)
SHABP (TP 2487)

- - - - ySHxB~D(g-qDjj
---T---7---- t T C W (DRDL)1

0.10-

- - - -k-----+--

0.00-

-0..20

0
I

-15.0

I
I
I
-10.0

I
I
I

I
-5.0

i
I

I
0.0

5.0

ANGLE OF ATTACK (deg)

Figure A-5. Normal force coemcient (wing-alone)at Mach 6.0

10.0

222 Aerodynamic Predictive Methods and their Validationin Hypersonic Flows

.---

-----I----

I -

MACH NO 6 . 0 ~ -

-7----7

I
EXPT ( ~ 2467)
h

I
I
I

-I
I

I
I

ANGLE OF ATTACK (deg)

Figure A-6. Pitching moment coemcient (wing-alone)at Mach 6.0

Appendix 223

I
I

I
I
- - - - - _I

I MACH NO 6.0 I
I
I

SHABP ( D R D ~ )

T P n V (DRDL) I
I

I
I

-6-7

ANGLE OF ATTACK (deg)

Figure A-7. Normal force coemcient (wing-body)at Mach 6.0

INDEX
A

Adiabatic wall 39
Aerodynamic
preliminary analysis II
(APAS II) 109
heating at hypersonic
speeds 67
performance 3
preliminary analysis II
(APAS II) 109
Aerodynamics 175
body-alone 34
wing alone 45
of rarefied gas flows 175
Ailerons 156
Analysis
drag buildup 96
APAS 97
code 156
APAS II 109
Aspect ratio 107
Asynchronous elevon deflection 150
Axial pressure coefficient 40

Circular cylinder 45
Compressibility factors 39
Compressible turbulent flow
62
Compression flow 46
Computation of
aerodynamic coefficients
98
Cone 16, 211
frustrum 195
Convective heating equation
83
Cylinder 42, 208,

Embedded Newtonian flow 8


Empirical equations 65
Energy transfer 203
Enthalpy 68, 73, 74
recovery 70
stagnation 70
empirical 65
Expansion 46

Base pressure coefficient 39


Bessel function 193
Bessel functions 193
Blast wave 18
Boundary layer transition
86

D
DahlemBuck empirical 109
method 17
Delta wing emperical
method 14
Direct simulation monte
carlo technique 176
Double circular arc 56

226

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

Isentropic expansion 11

Flow regimes 176


Fluid dynamic simulation
201
Force
axial 37, 190, 197
normal 41, 50, 188, 190
Free molecule flow 176, 203

Mach number 14, 97


Maxwellian
distribution function 183
equilibrium velocity distribution function 179
Method
DahlemBuck empirical 17
delta wing emperical 14
Hankey flat surface empirical 17
hypersonic impact 96
nonlinear CFD 168
OSU blunt body 17
panel 95
prediction method 95
Quinn & Gong 80
second-order shock-expansion 128, 138
Sommer & Short 62
Spalding & Chi 64,
98,108, 109
subsonic panel 156
tangent wedge, tangent
cone 14
tangent-cone empirical 96
tauber 87
Van driest-II 63
vortex lattice 95
MISLIFT 97
MNT 7
Modified
Newtonian theory 7
circular arc 56
double wedge 55
Monte carlo technique 200

Gentry program 107


Gentrys
hypersonic arbitrary body
aerodynamics 139
program (HABP) 109

H
HABP 101, 108, 168, 215
predictions 158
Hankey flat surface empirical
method 17
Heat flux 69
transfer analysis 80
transfer coefficient 75
transfer methodology of
Tauber 87
Heating
analysis 67
rates cone flat plate 89
Hemisphere 45
Hemispherical nose 42
HYFAC 101
Hypersonic
impact methods 96
isolation principle 107
Mach numbers 110
rarefied transitional flow
202
speeds 67
research airplane 95, 110,
139

I
Inclined cone 109
Inviscid zero lift drag 108

Knudsen number 175

L
Laminar flow 66, 78

Index

N
Newtonian & Prandtl-Meyer
model 10
Newtonian theory 6
North American X-15 110

O
OSU blunt body method 17

P
Pan air 97
Panel method 95
Pitching moment 52
coefficient predictions 149
characteristics 119
Planform area 107
Pointed
cone 40, 41, 44
ogive 42, 44
Potential flow theory 110
Prandtl number 71
Prandtl-Meyer expansion
108
flow 96
Prediction methods 95
Preliminary design analysis
158
Program
supersonic implicit marching (SIMP) 108

R
Ramp surface 9
Real time heating analysis
80
Recovery enthalpy 70
Reynolds
analogy factor 77
number 65
Rolling moment 149
Rudder 51

S
SHABP 215
Shuttle orbiter 109

SIMP 108
Single
parabolic arc 58
wedge 55
Skin friction
coefficient 37
forces 61
Slip & transitional
flow regime 212
flows 200
Solar radiation 68
Sommer-Short
estimate 108
skin friction estimate 108
Space shuttle orbiter 156
configuration 156
Spalding & Chi method 64
Sphere 194, 208
Spherical
nose 69
segment 198
Stagnation
density 71
enthalpy 70,74
line heat transfer 72
point 71
heat transfer 69,88
heating rate 80
temperature 71
Stanton number 77
Subsonic flow analysis 95
Sutherland law 71
Sutherland relation 79
Swept infinite cylinder 88
Swept wing 72

T
Tangent
cone 101
wedge 109
Tangential accommodation
coefficients 186
Taper ratio 107
Theory
2-D airfoil 25

227

228

Aerodynamic Predictive Methods and their Validation in Hypersonic Flows

theory in hypersonic
flows 25
Allen & Perkins viscous
cross flow 43
first order 18
linear 108
linear supersonic 96
modified Newtonian 7, 34
Newtonian 6
potential flow 110
Prandtl-Meyer 98
expansion 10
second order shock expansion 19, 168
SOSET 19
Shock expansion 18
tangent-cone 98
unified supersonic-hypersonic small disturbance
128
Van dyke unified 24
small disturbance 116
Thickness ratio 107
Transitional flow 176

Turbulent flow 66, 78

U
Unified distributed panel
109

V
Validation 95
Vortex lattice method 95

W
Walker and Wolowicz
115, 138
Wedge 16
Wing
cone configuration 107
leading edge variations
107
sections 52

X
X-15 118, 168

Y
Yawing moment 138, 149

You might also like