Sei sulla pagina 1di 524

Computational Methods in

Hypersonic Aerodynamics
Edited by

T.K.S. Murthy

Computational Mechanics Publications


SOUTHAMPTON / BOSTON

Kluwer Academic Publishers


DORDRECHT / BOSTON / LONDON

T.K.S. Murthy
Associate Director of Extension Programmes
Wessex Institute of Technology
Ashurst Lodge
Ashurst
Southampton S04 2AA
U.K.
Published by Kluwer Academic Publishers,
P.O. Box 17, 3300 AA Dordrecht, The Netherlands.
In co-publication with Computational Mechanics Publications
Ashurst Lodge, Ashurst, Southampton, U.K.
Sold and distributed in the U.s.A. and Canada
by Kluwer Academic Publishers Group,
101 Philip Drive, Norwell, MA 02061, U.S.A.
In all other countries, sold and distributed
by Kluwer Academic Publishers Group,
P.O. Box 322,3300 AH Dordrecht, The Netherlands.
British Library Cataloguing-in-Publication Data
A Catalogue record for this book is available
from the British Library
{1SBNXl-7923-1673-8 Kluwer Academic Publishers Dordrecht/Boston/London
'ls13N 1-85312-156-8 Computational Mechanics Publications, Southampton
ISBN 1-56252-083-0 Computational Mechanics Publications, Boston, USA
Library of Congress Catalog Card Number 91-77003
All rights reserved. No part of the material protected by this copyright notice may be reproduced
or utilized in any form or by any means, electronic or mechanical, including photocopying, recording
or by any information storage and retrieval system, without the written permission of the publisher.
@Kluwer Academic Publishers, 1991
@Computational Mechanics Publications, 1991
@See also p. 151
Acknowledgement is made to P. Gnoffo for the use of Figure 8 on p. 140 which appears on the front
cover of this book.
Printed and bound in Great Britain
The use of registered names, trademarks etc., in this publication does not imply, even in the absence
of a specific statement, that names are exempt from the relevant protective laws and regulations and
therefore free for general use.

COMPUTATIONAL METHODS IN HYPERSONIC AERODYNAMICS

CMP

.,
~

FLUID MECHANICS AND ITS APPLICATIONS


Volume 9
Series Editor:

R. MOREAU
MADYLAM
Ecole Nationale Superieure d' Hydrau/ique de Grenoble
Bofte Postale 95
38402 Saint Martin d'Heres Cedex, France

Aims and Scope of the Series


The purpose of this series is to focus on subjects in which fluid mechanics plays a
fundamental role.
As well as the more traditional applications of aeronautics, hydraulics, heat and
mass transfer etc., books will be published dealing with topics which are currently
in a state of rapid development, such as turbulence, suspensions and multiphase
fluids, super and hypersonic flows and numerical modelling techniques.
It is a widely held view that it is the interdisciplinary subjects that will receive
intense scientific attention, bringing them to the forefront of technological advancement. Fluids have the ability to transport matter and its properties as well as
transmit force, therefore fluid mechanics is a subject that is particulary open to
cross fertilisation with other sciences and disciplines of engineering. The subject of
fluid mechanics will be highly relevant in domains such as chemical, metallurgical,
biological and ecological engineering. This series is particularly open to such new
multidisciplinary domains.

The median level of presentation is the first year graduate student. Some texts are
monographs defining the current state of a field; others are accessible to final year
undergraduates; but essentially the emphasis is on readability and clarity.

For a list of related mechanics titles, see final pages.

CONTENTS
Preface

VB

Chapter 1: Introduction to the Physical Aspects


of Hypersonic Aerodynamics
R. Grundmann

Chapter 2: Computational Methods for Viscous


Hypersonic Flow
D. Hanel

29

Chapter 3: Numerical Simulation of Hypersonic Flows


J. S. Shang

81

Chapter 4: Point-Implicit Relaxation Strategies for


Viscous Hypersonic Flows
P. A. Gnoffo

115

Chapter 5: Flux-Split Algorithms for Hypersonic Flows


P. Cinnella, B. Grossman

153

Chapter 6: Efficient Multigrid Computation of


Steady Hypersonic Flows
B. [(oren, P. W. Hemker

203

Chapter 7: Laminar-Turbulent Transition


D. Arnal

233

Chapter 8: Second-Order Effects in Hypersonic


Laminar Boundary Layers
J. Ph. Brazier, B. A upoix, J. Cousteix

265

Chapter 9: Real Gas Effects in Two- and Three-Dimensional


Hypersonic, Laminar Boundary Layers
B. A upoix, J. Cousteix

293

Chapter 10: Flow Analysis and Design Optimization


Methods for Nozzle-Afterbody of a Hypersonic Vehicle
O. Baysal

341

Chapter 11: The Computation over Unstructured Grids


of Inviscid Hypersonic Reactive Flow by Upwind
Finite-Volume Schemes
J.-A. Desideri

387

Chapter 12: Computational Aerothermodynamics for


2D and 3D Space Vehicles
J. Hauser, J. Muylaert, H. Wong, W. Berry

447

Index

491

PREFACE
This book contains chapters written by some eminent scientists and researchers on
Computational Methods in Hypersonic Aerodynamics and is a natural sequel to the
earlier books on Computational Methods in Potential Aerodynamics (1986) and Computational Methods in Viscous Aerodynamics (1990) published by Computational
Mechanics Publications in their Aerodynamics Series.
In this book, the earlier attempts at the solution of the highly non-linear NavierStokes equations are extended to the aerothermodynamics of flow in the hypersonic
regime. It includes the effects of viscosity on the physical and chemical processes
of high-temperature non-equilibrium flow at very high speeds, such as vibrational
excitation, dissociation and recombination, ionization and radiation, as well as real
gas effects and the effects of high temperature and low density.
The opening chapter by Grundmann provides an introduction to hypersonic flow
with a description of the physical (and chemical) aspects of the flow. As an example, the author describes these processes involved in the flow along the flight path
of an orbital vehicle with a basic consideration of fluxes and transport properties.
Grundmann also provides a specimen of the non-equilibrium boundary layer along a
flat plate. The chapter concludes with a useful set of definitions for a reacting gas
mixture in high-temperature flow.
Computational methods for the progressive solution of the Navier-Stokes equations
are presented in the second chapter by Hanel. The governing equations and relations
for equilibrium flow are first described with an outline of their numerical formulation,
including conservative discretization and estimation of fluxes. The effect of numerical
damping on the solution is demonstrated for the steady-state solution. The chapter
concludes with a consideration of the numerical problems of real viscous, hypersonic
flows.
In Chapter 3, Shang continues the numerical simulation of hypersonic flow. This
author explains that the current methods of solution deal with continuum flow and
attempt the solution of the Navier-Stokes equations in conjunction with chemical
physics equations, including the non-equilibrium chemical reactions and internal degrees of excitation. In the rarefied gas domain, continuum concepts cannot be relied
upon and particle simulation is necessary. Shang discusses this briefly together with
the added mechanism of the energy transport process in radiation.
Chapter 4 by Gnoffo contains an upwind biased, point-implicit relaxation algorithm using the finite volume formulation for the numerical solution of the governing
equations for three dimensional, viscous, hypersonic flow in chemical and thermal
non-equilibrium. The author also includes an overview of physical models employed
for thermo-chemical non-equilibrium with several test cases and comparison with experimental data for hypersonic flow over blunt bodies.
Cinella in Chapter 5 reviews several numerical techniques for the simulation of fluid
flows spanning the range of reactive regimes, from local chemical equilibrium to full
thermo-chemical non-equilibrium. In particular, characteristic-based algorithms of the
flux vector and the flux-difference type are considered together with a discussion of
the problems associated with the modelling of thermo-chemical behaviour of reactive

mixtures of thermally perfect gases.


The next chapter by Koren and Hemker contains an efficient multi-grid computational method for steady, two-dimensional flow and for the hypersonic launch and
recovery flow around a blunt forebody with canopy. They consider solution of steady,
two-dimensional, Navier-Stokes equations by the upwind finite-volume method and
suggest improvements to the standard iteration and relaxation method. The robustness of the usual nonlinear multigrid method is improved by local damping of the
restricted defect, a global upwind prolongation of the correction and a global restriction of the defect.
Chapter 7 by Arnal is devoted to a description of theoretical, numerical and experimental problems related to boundary layers at high speeds. Starting with linear
stability theory, the author proceeds to the prediction of the location of transition by
the so-called "e" method and models the transition region by using, what he calls,
intermittency methods and other transition mechanisms. There is also a comparison
with experimental results.
The conventional boundary layer theory of Prandtl (extended by van Dyke) is unable to deal with vortical flows in the shock layer of hypersonic flows, as the matching
between the boundary layer flow and the external inviscid flow can never be perfect.
In order to improve this matching and to ensure a smooth merging of the viscous and
inviscid flows, Brazier, Aupoix and Cousteix (Chapter 8) propose a defect approach
coupled with asymptotic expansions. They also give some calculations on simple twoand three-dimensional bodies in hypersonic flow and compare with solutions of the
complete N avier-Stokes equations.
In the next chapter (Chapter 9) Aupoix and Cousteix again discuss in detail hypersonic, laminar boundary layers, with the emphasis on real gas effects which deeply
affect the evolution of the flow. However, thermal non-equilibrium effects are neglected
and attention is confined to chemical non-equilibrium effects. Self similar solutions are
also given for the flow and the wall heat flux at two- and three-dimensional stagnation
points or along attachment lines.
Baysal in Chapter 10 analyses the flow around the nozzle-afterody of a hypersonic
vehicle and suggests some design optimization methods.
In Chapter 11, Desideri also discusses the computation of inviscid, hypersonic,
reactive flows, using unstructured grids in upwind, finite-volume schemes. Equilibrium
flow solutions are obtained by combining the basic method (implicit, time integration
algorithms, employing local timestepping) with a classical algebraic model solved by
Newton method. Solutions to recently proposed non-equilibrium chemistry models
are found by integrating forward in time additional species of convection equations
with source terms.
In the final chapter, Hauser, Muylaert, Wong and Berry discuss computational
aerothermodynamics for 2D and 3D space vehicles. They start with an overview of
current and planned hypersonic activities of the European Space Agency. The special
physical aspects of high speed flows are then discussed, leading to the derivation of the
scaling laws and the relevant parameters characterising perfect gas equilibrium and
non-equilibrium flows. This then leads to a discussion of the mathematical formulation
of the governing physical equations and the physical submodels. Numerical solution
techniques are presented for multigrid Navier-Stokes equations emphasising the need

for flux linearisation. The authors conclude with a section on grid generation and
provide some examples of advanced visualisation of hypersonic flow.
The book has been prepared as a valuable contribution to the state-of-the-art on
computational methods in hypersonic aerodynamics. All the chapters have been written by eminent scientists and researchers well known for their work in this field. The
Editor wishes to express his special thanks to them for their cooperation during the
various stages leading up to actual publication.

The Editor
Southampton, January 1992

Chapter 1:
Introduction to the Physical Aspects of
Hypersonic Aerodynamics
R. Grundmann
Von Karman Institute for Fluid Dynamics,
B-1640 Rhode-St-Genese, Belgium
1. INTRODUCTORY REMARKS
Any obstacle, like meteorites, satellites or manned space vehicles, coming from space
and approaching the earth's atmosphere is subjected to physical flow phenomena.
These change its state of motion and also the state of its surface.
The surface temperature for example will be influenced. The friction caused by the
increasing number of air molecules the closer the obstacle approaches the ground heats
the surface by conduction, convection, radiation and other heat transfer mechanisms.
All this coincides with changes in the surrounding air becoming denser and denser. A
bow shock will appear in front of the object. A region with very high temperatures
around the body's nose will develop which has the energy to split even molecules
like oxygen or nitrogen into their atoms. This gives the opportunity to form new
combinations like nitric-oxygen. Single electrons may even leave the structure of an
atom or a molecule and the air becomes a partially ionized plasma.
One of the most important topics of research in hypersonic aerodynamics is to
find a reasonable way of calculating the temperature which is felt by a surface. Any
blunt nosed reentry body experiences at a velocity of about 11.2 km/s. It is the
orbital escape velocity. The temperature behind the bow shock belonging to this
velocity range is about 11 000 K. For comparison the temperature of the sun's surface
is about 5 000 K.
A temperature of 11 000 K will have an important feed back on the chemical
constitution of the air. Physical and chemical processes like vibration, dissociation,
recombination, ionization, radiation and many others arise. They need to be formulated to describe the phenomena involved in flows of these very high speeds and
temperatures. Simplified methods of calculating the temperature in the nose region
of the reentry vehicle will fail.
Perfect gas in inviscid flow
To demonstrate this, the total temperature To first will be calculated for a calorically
perfect gas in an inviscid flow. This total temperature is computed from the integral
energy equation assuming that the kinetic energy u~/2 of the flow is decelerated to
zero and completely transferred into a total temperature rise To. The total enthalpy
Ho is then written:

2 Computational Methods in Hypersonic Aerodynamics

Ho = cpTo = cpToo

u2

+ :;

For a Mach number of 31.6 which corresponds to the velocity of 8910 m/sec and an
altitude of 78 km above the surface of the earth the total temperature To will read:

To
with

,-1

= Too(1 + -2-M!) = 39540J{

Too = 197K
Moo = 31.6
, = 1.4
h = 78km

where Too is the static temperature of the atmosphere at the altitude h of 78 km. The
ratio of the specific heats, for a calorically perfect gas is constant.
The calculated result of approximately 40.000 K is much too high. This is because
some physical effects in high velocity flows were neglected. The existence of a bow
shock in front of the blunt body was not considered in the calculation. Behind a shock
the velocity is less than in front of it and therefore the kinetic energy is less which is
expressed in the energy equation.
The application of shock relations again for a calorically perfect gas at very high
Mach numbers will be used now for the static temperature Tshock behind a normal
shock. The shock is assumed to be strong, namely normal, which corresponds to a
shock angle f3 of 900 :
Tshock

- 1) 2 2 f3
= Too 2,(,
(
,+1 )2 Moo sm = 38250K

This temperature is already smaller by a considerable amount (1290 K), but still far
from being correct. The total temperature TO,shock has the same value as the previously
calculated total temperature To in front of the shock, since in an adiabatic perfect
gas the total enthalpy is constant across the normal shock. Hence, these formulas for
inviscid compressible flow computations are not applicable to these types of hypersonic flows. One of the reasons is that the above computed temperatures are purely
translational temperatures since the air particles only perform translational motions.
At high temperatures other energy modes may be excited like the rotational and vibrational modes for molecules. The molecules also perform rotations and vibrational
motions. These modes will reduce the calculated values for the pure translational
mode.
Another reason is that the viscous effects are missing which make the temperature
vary within the boundary layer. Its maximum is not necessarily located at the surface
of the body but close to it.
Furthermore, energy is transported in different modes for which exist different
energy equations. Energy is consumed for dissociation processes and many other
reasons are responsible for measuring a much lower temperature than calculated with
simple formulas given above. Additionally, the gas was assumed to be calorically
perfect which is not true at very high temperatures.

Computational Methods in Hypersonic Aerodynamics 3

A thermally perfect gas gives an even better result, since there the specific heats
and their ratio, are now functions of the temperature.

It must be imagined that chemically reacting flows need even more complicated
definitions of these properties, as will be seen later. The ratio of the specific heats,
will additionally depend on the pressure p.
, = ,(T,p)

From these considerations it is obvious that a very careful modeling of the physics
of hypersonic flows has to be performed. Otherwise the expected good results of numerical approaches are too far away from reality. The following sections are meant to
give an introduction to hypersonic aerodynamics. The physical phenomena of thermochemical non-equilibrium flows in this high temperature environment are described.
The glide path of an orbital vehicle leading through this environment of various energy
excitation zones is discussed. These physical effects have to be modeled in the governing set of the fluid mechanical equations. For this purpose some basic considerations
on fluxes and transport properties in hypersonic flows are given. Sample computations of chemical non-equilibrium boundary layers along a flat plate will show the
general behavior of high temperature effects. Finally, a set of definitions for reacting
gas mixtures in high temperature flows will be provided to lead into the kernel of this
book on computational methods in hypersonic aerodynamics.
2. DESCRIPTION OF THE FLOW PHENOMENA
Between subsonic and supersonic flows there is a decisive border which presents itself
physically by a shock, namely, a jump in pressure. This shock is clearly audible when
a flight object accelerates from subsonic to supersonic speeds.
The non-dimensional number describing the ratio of the local velocity u to the
local speed of sound a of a calorically perfect gas is called the Mach number M. It is
unity in case of the transient from subsonic to supersonic velocities.
u
M=a

<=> Mach number

where

, = V,RT
a

R
T

<=>
<=>
<=>
<=>

local speed of sound


ratio of specific heats
specific gas constant for air
local static temperature

A distribution of the temperature T and the local speed of sound for a calorically
perfect gaS is shown in Figure 1.

4 Computational Methods in Hypersonic Aerodynamics

90

80
altitude h
(kml
62
52
47

J119sosphere ...

32

20
11

I....

..

,,

stratosphere

aircraft

troposphere

clouds

0
0

50

100

150

200

250

300

350

T [K 1. a [m/s 1

Figure 1: Atmospheric temperature and speed of sound distribution versus the altitude

The temperature versus altitude distribution is based on the U.S. Standard Atmosphere from 1962 given in [1]. All free stream conditions in the following figures
and calculations are referring to these tables.
The earth's atmosphere consists of two main layers as explained in Fig. 1. The
first one is called homosphere and its extension is from the sea level to 90 km. Beyond
this the layer is called heterosphere. In the homosphere the composition of the air
is almost constant. There is about 78%N2 present and 21 %0 2, the remaining 1%
consists of other gases like argon and carbon dioxide. Beyond the homosphere the
composition of the air changes accordingly with the very high temperatures existing
there. Dissociation and ionization effects occur, which change the composition of the
air by creating for example nitric-oxygen and other products.
The homosphere is again split into three major parts which are the troposphere
extending from 0 to 11 km, the stratosphere extending from 11 to 20 km and the
mesosphere from 20 to 90 km. The troposphere contains 75% of the mass of the
atmosphere. A further 20% of the mass is in the stratosphere. That is, 95% of the
mass of the atmosphere is located within a layer of 20 km extension above the surface
of the earth.
In the eight shells in the homosphere it is known that the temperature changes
nearly linearly with altitude, see Fig. 1. In the heterosphere the temperature increases
linearly as well, but the gas no longer remains a continuum because the distance
between single molecules becomes larger and larger.
The curves in Fig. 1 are given for the atmospheric static temperature T and the
speed of sound a at zero motion of the airflow. In hypersonic flows the speed of sound
cannot be calculated applying the calorically perfect gas assumption anymore. Due
to the very high temperatures, the ratio of the specific heats now becomes a function

Computational Methods in Hypersonic Aerodynamics 5


of the temperature as well. In that case the gas is thermally perfect.
It should be mentioned here that the formula given above for the local Mach
number is not valid for the calculation of the equilibrium speed of sound for chemically
reacting mixtures. The general formula valid also for chemically reacting gas mixtures
at high temperatures is:
a

= J(8p/8p).

{:} local speed of sound at constant entropy

where p is the pressure and p is the density. The suffix s stands for isentropic flow
conditions, meaning that within a sound wave the flow conditions are reversible and
adiabatic. All thermodynamic properties of chemically reacting flows are functions of
the temperature and the pressure. The partial derivative of the pressure with respect
to the density may be replaced, such that the result for the equilibrium local speed
of sound becomes:

RT[l+(~)(fvh]

2_

(Oh)

a - I

[1 - P op T]

with
I

= cCp = f(T,p)
v

where e denotes the inner energy, v the volume and h the enthalpy for an equilibrium
chemically reacting gas. These thermodynamic quantities are combined through the
first and second law of thermodynamics and describe an isentropic process of a sound
wave. If the gas is calorically perfect, the formula reduces to the earlier one as expected. In that case the flow is frozen meaning the chemical state of the fluid does
not change with temperature.
Looking for a criterion to describe the transient from supersonic to hypersonic
velocities is not obvious. There is no visible or audible border as for the transient
from subsonic to supersonic flows.
A convention says that the border is placed around the Mach number:
M

{:} onset of hypersonic flows

This border is not fixed to that certain number, but around M ~ 5 some physical
phenomena become more and more important and therefore characterize hypersonic
flows. In the following text some of these conceptions, frequently mentioned in the
literature, are presented and discussed, see also ANDERSON [2,3].
Due to the very high speeds, the kinetic energy of the flow is considerable. When
this flow is decelerated by friction effects close to body surfaces an energy transfer to
thermal energies is observed. These effects are usually called strong viscous effects.
The boundary layer becomes thicker now depending also on the Mach number of the
flow.

6 Computational Methods in Hypersonic Aerodynamics

<=> boundary layer thickness in hypersonic flows


In this equation Re is the Reynolds number defined as follows:
puL

Re=--

<=> Reynolds number

J-l

where p is the density, u is the velocity, L is a characteristic length of the vehicle,


and J-l is the dynamic viscosity. In general the Reynolds number is a high number in
subsonic and supersonic flows. At high altitudes the density p is low, so becomes the
Reynolds number. Including a high Mach number this together gives a much thicker
boundary layer than in flow regimes with lower speeds.
The boundary layer increases and this is felt by the surrounding inviscid flow
field, which again influences back on the viscous part. This feed back is called viscous
interaction. The boundary layer can become so thick that it is of the order of the thin
shock layer itself which is located between the shock and the body surface. In this
case the layer is called fully viscous shock layer.
Within this layer the concept of boundary layer theory is no longer valid because
of this strong interaction with the shock. The shock itself is no longer a very narrow
region of a discontinuity but it is smeared out and completely merged in the viscous
layer.
A flow that crosses a bow shock in front of a blunt body experiences on different
stream lines different entropy increases. The layer behind the shock is then called
entropy layer. In this layer the boundary layer will develop such that its outer boundary is still merged in the strong entropy gradients of this entropy layer. A problem is
immediately posed because of the outer boundary conditions for the boundary layer.
Outside of it, there are strong gradients of the flow properties. This is normally not
the case for boundary layer flows. Since the entropy layer is coupled with strong
vorticity, it is also called vorticity interaction.
Hypersonic flows are high energy flows. Kinetic energy will be transferred to
random thermal energy while the flow passes a shock, for example. This says that
the resulting temperature increase can be so large that the gas behaves in a non-ideal
manner.
The vibrational energy of the molecules becomes excited which consumes some of
the available thermal energy. For air this effect starts at a temperature of 800 K.
So, besides the translational and rotational energy of the molecules another mode
of thermal energy exists. The vibrational temperature TYib belonging to this kind of
energy can change the niveau of the translational temperature T.
Chemical reactions can occur when the temperature increases further. For an
equilibrium chemically reacting gas the specific heats become functions of the temperature and the pressure as well. Oxygen dissociates at 1 atm pressure at about 2000
K, while nitrogen begins to dissociate at about 4000 K. At temperatures above 9000
K ions are formed and the gas becomes a partially ionized plasma.

Computational Methods in Hypersonic Aerodynamics 7


An altitude versus velocity plot, Figure 2, includes these above mentioned gas
effects. The figure shows the constituent species in equilibrium air after passing a
normal shock, see also HANKEY [5].
If vibrational excitation and chemical reactions take place very rapidly, much
faster than a fluid particle moves in the flow, there is vibrational and chemical equilibrium. If the other way is true, then the flow is in non-equilibrium.
All these preceding energy excitation phenomena are called high temperature effects. Frequently they are called in literature real gas effects but this is a misnomer
since real gas effects include intermolecular forces. Normally these are negligible in
fluid dynamical considerations of gas flows.
At low densities the smearing out of the shock becomes worse and PRANDTL's
no-slip condition at the wall also no longer holds. The velocity at the wall is not zero
any more but finite. The boundary condition at the wall is then called slip condition.
Comparable things happen to the temperature, the temperature slip condition is now
valid.
These effects at low density are called low density effects. If the density becomes
so low that the mean free path of the molecules becomes of the same order of the
characteristic length of the body, the flow will be no longer a continuum. Then it
becomes a so-called free molecular flow which is usually treated by kinetic theory
techniques.
All these phenomena described above are characteristics of hypersonic flows. They
may appear together or only one of them is dominating. Each flow situation has to
be considered separately and the characteristics have to be discussed. Order of magnitude analysis and dynamic force considerations may give hints which part is of
importance and which one is not. In the following, the flight pass of an orbital vehicle
will be designed. This leads through the different excitation zones of the different
gases existing in air given in Fig. 2.
3. FLIGHT PATH OF ORBITAL VEHICLES
A hypersonic vehicle having solved its mission in space reenters the atmosphere of
the earth to start its way down to the space port. For this purpose a certain glide
path has to be chosen to avoid structural damage of the fuselage or overheating of
the surface. On the other hand the descent should not last too long since another pay
load mission is envisaged.
Figure 3 presents an altitude versus velocity plot where such a glide path is given.
The lines indicate constant weight over lift times planform areas taken from HANKEY
[5]. The resultant density versus velocity formula develops from the equations of
motion for planar flight, see ANDERSON [4J:

V
L Fnonnal = mRearth
-- = L 2

where

{:} velocity
{:} lift

W cos ()

8 Computational Methods in Hypersonic Aerodynamics

10% 1P~'b% 90~ 0%

90%

10%

10

11

12

13

14

15

velocity [ km/s I

Figure 2: Energy excitation zones for the stagnation point

10

11

12

13

velocity [ km/s I

Figure 3: Reentry corridor in the altitude versus velocity plot

14

15

Computational Methods in Hypersonic Aerodynamics 9


W=mg

()=o

:>
:>

weight
incidence

In this formula EFnonnal lists all the forces directed normal to the flight path.
If the angle of incidence () is zero which may be assumed for a gradually horizontal
glide path, then the above mentioned formula solves for the density distribution p for
variable velocity V.

(~ _ ~)

p(h) = 2W

SCI

V2

Vc2

where W is the weight, S is the planform and CI is the lift coefficient of a shuttle type
aircraft. The velocity Vc is related to the gravitation g being also a function of the
altitude and the radius Rearth of the earth by the following formula:

v:,2 =

g Rearth

Since the density p is only a function of the altitude h the flight corridor results
in a band of the above mentioned constant load factor 2W/(Scl). This is a parameter
which lies between too high or too hot values.
In the same Fig. 3 constant free stream unit Reynolds numbers are also given.
Unit Reynolds number means that the Reynolds number is divided by a reference
length x of the vehicle. The formula describing this number is:
Rex

--;- =

p(h)V
J1(T)

This formula again may be solved for the density p which simply is a function of
the altitude h. The same yields for the viscosity J1 in the standard atmosphere.
p( h) = Rex J1(T)

x V
These constant unit Reynolds number curves indicate that the glide path leads from
pure laminar flow at high altitudes where the density is still low through transitional
to fully turbulent flow regions close to the surface of the earth. The horizontal lines
show the Reynolds number starting in the turbulent region at values of 3 . 108 down
to 3 10 in the laminar part.
The last set of curves in Fig. 3 are the free stream Mach numbers. They are based
on the assumption of a calorically perfect gas where the ratio of the specific heats
I is a constant, I = 1.4. The Mach number lines are shown from 8 to 48. They
are not constant with increasing altitude because they depend on the temperature
distribution, see Fig. 1.
These glide paths, also called reentry trajectories, lead from free molecular flow
to continuum flow. On the way they pass the physical phenomena of vibrational
excitation, dissociation of oxygen and nitrogen and even ionization. In Figure 4 the
current glide path is plotted including the various energy excitation zones for the
stagnation point.

10 Computational Methods in Hypersonic Aerodynamics

10%

10"10 90%
90%
10%

90%

10%

10

11

12

13

14

15

velocity [ km/s I

Figure 4: Energy excitation zones for the stagnation point

From this diagram it is seen that the region is very small where the air obeys the
ideal gas concept.
p=pRT

ideal gas equation

The remaining regions where the molecules are vibration ally excited, the oxygen
and the nitrogen begin to dissociate, and where ionization takes place have to be
described by the additional knowledge of the microscopic states of the various energy
levels. The gas considered in these cases is assumed to stay a perfect gas but now
consisting of different species. Now, it is a mixture of perfect gases.
~

ideal gas mixture equation

A region where the flow is in vibrational excitation starts very early in the plot.
The excitation begins at temperatures of 800 K and is terminated at about 2000 K.
Here the dissociation of oxygen begins. At around 4000 K it is completed and the
first nitrogen molecules split into their atoms. This process is finished close to the
9000 K temperature line.
The phenomenon of ionization, where atoms set electrons free and a plasma develops, begins at the same temperature border of about 9000 K.
All these regions are given in a band width from 10% to 90%. From the plot
it is to imagine that these regions partially overlap. Thus, a complete numerical or
experimental consideration of the flow field along the reentry trajectory of a space
shuttle must include all these flow phenomena. The complexity can be demonstrated

Computational Methods in Hypersonic Aerodynamics 11


on behalf of the total energy of a flow around a vehicle along this glide path. It
consists namely of the internal energies of all species, their kinetic energies, the sum
on all the vibrational energies, the electron translation energy, the sum on all species
electronic energies and the sum on all the zero-point or formation energies of this gas
mixture under consideration.
To count for all these energies and properties in thermo-chemical non-equilibrium
gas mixture flows the equations of change have to be prepared accordingly. These are
the momentum transport equations consisting of three spatial equations, the mass
transport equations for n species, the vibrational energy transport equations for m
molecular species, the electron translation energy transport equations for p ionizing
species and the global energy equation where radiative heat transfer must be included.
4. SOME BASIC CONSIDERATIONS OF FLUXES AND TRANSPORT
PROPERTIES
Before starting with the application of the equations of change for mass, energy and
momentum some similarities in their structure should be pointed out.
The equations of change are normally given in terms of fluxes corresponding to the
conservation law form. These fluxes are the mass flux by ordinary diffusion, the heat
flux by conduction and the momentum flux by molecular transport. To obtain concentration, temperature and velocity profiles indicating the primitive variables some
fluxes must be replaced by expressions that involve the transport properties and the
gradients of the concentration, the temperature and the velocity. The corresponding transport properties are the diffusivity, represented in Fick's law, the thermal
conductivity in Fourier's law and the viscosity, represented in Newton's law.
The mass flux does not only depend on the concentration gradient describing
the ordinary diffusion. Although this is a major effect, there are, even in isothermal
systems, three mechanical driving forces that cause a movement of a species in a bulk
fluid with a velocity of its own. These driving forces are the concentration gradient,
the pressure gradient and eventual external force differences.
In a multi-component system there are fluxes of mass, energy and momentum.
By thermodynamics of irreversible processes there will be a contribution to each flux
belonging to each driving force in the system. They may also be coupled from one
flux to the other.
The following table adopted from BIRD, STEWART and LIGHTFOOT [6J will
help to recognize the relationships between the driving forces and their corresponding
fluxes.
The mass fluxes due to a pressure gradient and due to the external force differences, are not listed in the table. The energy fluxes given by interdiffusion effects and
radiation are neglected here as well.
The hints that the momentum flux involves a tensor of second order and the energy and the mass fluxes are described by vectors show that a coupling can occur
between the force and flux relationship only when the tensor is of the same order.
For the energy and the mass fluxes and their forcing functions a coupling exists since
each gradient forces another flux to appear aside the diagonal shown in the table.
Consequently, in multi-component systems the mass flux depends both on the me-

12 Computational Methods in Hypersonic Aerodynamics

driving
force
fluxes

Concentration
Gradient

Mass
(first ord.
tensor=
vector)

Fick's
law

Energ~

Oufour
effect
OT

(first ord.
tensor=
vector)

Teml2erature
Gradient

Gradient

Soret
effect
OT

i j

Velocit~

Fourier's
law
k

Newton's
law

Momentum
(second ord.
tensor)

/1

Table 1: Driving forces and resultant fluxes

chanical driving forces, namely on the ordinary, pressure, and forced diffusion, and
on the temperature gradient, also called thermal-diffusion or Soret effect. The energy
flux depends both on the temperature gradient causing heat conduction and on the
concentration gradient which is either called diffusion-thermo or Dufour effect. The
momentum flux is dependent only on the velocity gradient.
In order to describe the Soret and the Dufour effect an additional transport property to the existing ones like viscosity /-l, thermal conductivity k and diffusivity Dij
has to be added. It is the thermal diffusion or the Soret coefficient Dr.
From the first two rows of the table it turns out that the total mass and energy
fluxes are sums of two elementary fluxes each. This is true for the considered case.
There are even more of these fluxes due to other mechanical driving forces not listed
in the table, like the pressure gradient or external force differences. Provisionally,
these existing mass fluxes are summed up.

The index s stands for the individual species s of the gas mixture. Since the mass
flux j. is a vector, another index should be added to precise its spatial extensions.
The so-called ordinary diffusion j~rdinary is the portion of the mass flux which originally depends on the gradient of the concentration of each species. The distribution
of the concentration of the species will be the result of the mass transport equations

Computational Methods in Hypersonic Aerodynamics 13


for each single species. These equations can also be understood as partial continuity
equations, since the sum on all species concentration equations finally yield the global
continuity equation.
The second term describes the mass flux caused by pressure diffusion j;ressure.
Assuming that a pressure field is superimposed on the flow field, then it causes a
movement of the species. Such a pressure field could appear under centrifugally
accelerated movements. Normally a flux due to imposed pressure is really small such
that it is negligible.
The third term presents the forced diffusion j;orced. The force can be the gravity
force, forces due to an electric field exerted to electrons or ions in the chemically
reacting flow, or any other external force.
The fourth portion of the flux j!hennal is the one which occurs due to a temperature gradient. The species are transported by this gradient because of their different
characteristic temperature properties. This flux is also called thermal diffusion. The
thermal diffusion is normally very small. It considers the quality that species try to
diffuse through others by a temperature gradient. This type of diffusion is also known
as the Soret effect.
Secondly, the existing heat fluxes are summed up.
q=

qconduction

+ L q?ufour + L q~nterdiffusion + tadiation


s

The internal summations on s in that equation count the energy fluxes due to the
diffusion-thermo effect and interdiffusion of each single species s. The heat flux q also
is a vector and therefore it is dependent on the spatial conditions.
The conductive energy flux qconduction is the dominant one for lower temperature
ranges. It is comparable to the ordinary diffusion term in the considerations on the
mass transport, since it is only dependent on the temperature gradient.
Due to the concentration of species, a heat transfer q~ufour is created which is
normally of negligible size compared to the others. It is also called Dufour or thermodiffusion effect and for completeness it is given in the second term.
The third heat flux term q~nterdiffusion is caused by the mass diffusion of the species
themselves. Therefore, its name is interdiffusion. Due to the different species mass
fluxes a transport of energy by the species is performed. Each species carries its own
enthalpy including its energy of formation which might be very different in chemically
reacting gases.
Last but not least, there is the radiative flux ifadiation The threshold temperature
at which air radiates is 10 000 K. Therefore, this term is of minor order in general.
However, the heat transfer by radiation should not be neglected when the thermal
load of a surface must be estimated correctly.
5. SAMPLES OF CHEMICAL NON-EQUILIBRIUM BOUNDARY LAYERS
ALONG FLAT PLATES
To illustrate some of the different fluxes given in the table above a Couette-flow is
considered in Figure 5. This is a laminar flow between two flat plates involving no
downstream pressure gradient. The upper plate moves from the left to the right with

14 Computational Methods in Hypersonic Aerodynamics

soluble wall moves

U=v

T=Tw
,,

flow direction

,,

P=P w

,
\

U=O

"

CA = CAmax

= Twad

:
:
:
:

P=Pwad

I
1
I

,
CA=

Figure 5: Couette-flow with velocity, temperature, density and concentration profiles

the velocity Voo. By doing that it forces the flow to move producing a linear velocity
profile as shown in the left part of Fig. 5.
Due to friction, a temperature and a density profile of the base flow develops. The
temperature shows a normal tangent at the non-moving wall which defines the state of
an adiabatic wall. No heat transfer by conduction is possible through this wall. The
wall takes the temperature T of the flow adjacent to it. Since there exists no normal
pressure gradient in y-direction, the density profile p follows the equation of state and
presents an opposite shape to the temperature profile. At high temperatures is the
density low. A specialty of the upper wall is that it is slightly soluble meaning that
a very little amount of it will diffuse into the fluid flow. The amount is so small that
the mass balance is not disturbed. Therefore the mass fraction of the wall material PA
has its maximum at the upper moving wall while the concentration decreases slowly
in the negative y-direction approaching the value zero.
For the friction in the flow stands the transport of momentum expressed by Newton's law.

au

Txy

= -/1 ay

{:} shear stress

The shear stress Txy is the only momentum flux in the equation of change for the
momentum. It can be influenced by other transport mechanisms only via the viscosity
/1, since this is a function of the temperature. The momentum transport takes place
always in the direction of decreasing velocity.
The energy transport by conduction describes a heat flux. The temperature gradient is the means to transport energy in the direction of decreasing temperature.
Fourier's law of heat transfer by conduction is:

Computational Methods in Hypersonic Aerodynamics 15


{)T

qy =-k{)y

<=> heat transfer by conduction

The material of the soluble wall will be transported in the negative direction to
the wall again in the direction of decreasing mass fraction PA. The law to model this
movement was given by Fick. It is the so-called law of diffusion.

JA

= -

D AB-{)PA

<=> mass transfer by ordinary diffusion

{)y

As stated above another mechanism of mass transport is caused by a temperature


gradient. This effect is called Soret or thermal diffusion effect.

<=> mass transfer by thermal diffusion


For the Couette flow in Fig. 5 there are two kinds of mass fluxes involved. Together, they are the basic transport mechanism for the heat transfer due to a mass
concentration. This kind of heat transfer is called Dufour or diffusion-thermo effect.
.

qAy

= ( JAy

.T )

+ JAy

PA

(1

'RTI
-

PA

D~

1MA )-D
AB

<=> heat transfer by diffusion-thermo effect


Now, to complete the heat transfer calculation the heat transfers by conduction
and by the Dufour effect have to be added.
Plotting these calculated mass and heat fluxes given by the preceding formulas
will show the same shapes of the basic variables like density and temperature but
with an altered sign. This says that the energy flux tends to lower temperatures and
the mass flux to lower concentration of the soluble species PA.
In the following section some examples of a chemically reacting non-equilibrium
boundary layer in a diatomic gas are presented including the above described transport
mechanisms.
Non-equilibrium boundary layers along flat plates
Some examples of two-dimensional boundary layer calculations in a hypersonic flight
regime are presented here for an illustrative introduction to chemically reacting flows.
Temperature and pressure dependent molecule and atom distributions will be shown
and discussed in the text in a qualitative way.
The flow is supposed to be in thermal equilibrium, meaning that the thermal
time scales are much smaller than the chemical time scales. This is true for a couple
of thousand degrees of temperature. The chemical processes are computed for nonequilibrium but during the computations they may reach their equilibrium state.
The reason for the different time scales is that all existing energy modes are
quantized. The translational, the rotational, the vibrational and the electron energy
modes are not changing continuously in level but microscopically seen in discrete

16 Computational Methods in Hypersonic Aerodynamics

5.-------,-------,
x = 0.0 m
0.1 m
4 1Ii--... -.. -..-..-..-..-.-t-=-1-:-.-=-0-m---i
X

x=

x = B.O m

y/6

4 ~4--------~--------~
,~

y/6

31---\~---f------__1

.---------,---------~

~.~

\~

~~---~----~

.....\

\'.. \

2 r--.~~.~~..~-~-----~

'\:...., ,

1f-----':-+--1'-:----'~---1

~ "-

~------~~~----~~

25

50

Figure 6: Temperature and concentration profiles along an adiabatic flat plate (Ma

= 15)

steps. These steps are performed by particle collisions. It takes several collisions to
change the level in the different energy modes. The number of collisions needed for
thermal equilibrium, about 20 000, is ten times less than for chemical equilibrium. If
now the time scale of thermal equilibrium is less than the one of chemical equilibrium,
then the energy mode of the vibrational energy is already in equilibrium before the
chemical process is finished.
The gas under consideration is a diatomic gas like oxygen or nitrogen. It only
consists of a concentration of atoms and molecules. A mixture of oxygen and nitrogen
is not considered here.
Vibrational energy modes, electron translation energies, ionization and radiation
are excluded. Because a boundary layer flow is considered here, the effect of catalycity
at the wall on the heat transfer can be discussed. The reactions with the wall material
are called heterogeneous reactions while inside the flow homogeneous reactions take
place. This difference will be pointed out by the sample calculations.
The geometry of the flow problem is simply a flat plate which excludes surface
curvature effects and the downstream pressure gradient is assumed to be zero. The
velocity profiles in a boundary layer along a flat plate are similar. They do not change
in shape but in size with increasing boundary layer thickness. Therefore, they will
not be shown here.
Examples of this two-dimensional chemically reacting non-equilibrium boundary
layer flow for a diatomic gas were given by PETERS [6]. Temperature and concentration profiles of atomic oxygen are shown for increasing downstream distance in Figure

6.

In both parts of the figure y 18 is the non-dimensional distance normal to the


surface where 8 is defined as the boundary layer thickness. The non-dimensional
quantities along the abscissa are the temperature T ITe and the atomic oxygen concentration CAl CAe. The suffix e denotes the value at the outer boundary layer edge

Computational Methods in Hypersonic Aerodynamics 17


used for non-dimensionalization. In all sample computations the temperature Te at
the boundary layer edge is approximately Te = 200 K.
The maximum of the temperature profiles is located at the wall. The inner energy
generated by dissipation can not penetrate the adiabatic surface which leads to a
temperature increase of about 10 000 K. This causes dissociation of the molecular
oxygen O2 into atomic oxygen 0 by consuming inner energy. Thus, a decrease of the
wall temperature follows with increasing downstream direction while the concentration
of the oxygen atoms rises. Due to the exponential temperature dependency of the
production density high temperatures are reduced stronger which causes a flattening
of the temperature profiles.
To compute a flow with catalytic chemical reactions between the wall material and
the flow species an appropriate wall boundary condition for the species concentration
equation must be specified.
The chemical reaction at the wall is governed by the catalytic behavior of the
wall material. Glass-like surfaces will influence the reactions differently from surfaces
made of metal. The mass flux at the wall has to be considered in balance between
the catalytic wall material and the gas mixture next to it, reminding the definition of
the mass fluxes prepared already in the preceding section.

. +'T _

or

catalytic

JAw JAw - JAw


8PA
T 1 8T
-DAB 8y Iw-DATw8y Iw=-kwPAw

The first term on the left hand side describes the ordinary diffusion at the wall and
the second one involves the mass flux based on a temperature gradient also known as
the thermal diffusion or Soret effect.
The right hand side is simplified. The knowledge about the reaction rate of the
wall material with the gas mixture is subsumized in the empirical rate coefficient kw.
The density PA of the oxygen atoms can be transferred to the concentration CA
by simple transformations and definitions which will be reported later.
This boundary condition of the so-called third kind is one of the very rare cases. It
involves the value of the concentration CA and its gradient in normal direction 8CA/ 8y,
prescribing together a Dirichlet and a von Neumann boundary boundary condition.
Additionally, the temperature gradient at the wall finds a place in this complex mass
flux boundary condition.
From this boundary condition of the third kind it results that a normal gradient of
the concentration profile at the wall can be generated by two reasons: the wall behaves
adiabatic and the thermal diffusion or Soret coefficient D~ is set to zero because it is
negligible. In both cases the wall catalycity kw must have the value zero.
Figure 7 shows another set of temperature and atom concentration plots for this
second case of mass flux boundary conditions. Prescribing a constant wall temperature
of about 660 K changes the boundary condition for the concentration also. The
catalycity of the wall and the thermal diffusion coefficient were assumed to be zero.
The maximum of the temperature profiles reduces while going in downstream
direction, since the energy is consumed by dissociation. The boundary condition
for the concentration prescribes a normal gradient of zero. No atoms are forced to

18 Computational Methods in Hypersonic Aerodynamics

kw=
4

'W=

0 m/sec

x= Om
x= 2m
x= 10m
x=200m

y/6
3

'.~

'. ..~
, '. ........
'

0
0

15

TfTe

30

..

."\

y/6

0 m/sec

,,

"

....... .....
,,

.....
"-

"-

"-

......
]
,;

:."
/.

,""
0

0.10

."

e Ie 0.20
'A Ae

Figure 7: Temperature and concentration profiles at constant wall temperature along a flat plate
(Ma = 24)

recombine due to the presence of a catalytic wall material. The concentration increases
at first in downstream direction. The temperature decreases far downstream because
of the growing boundary layer thickness. Thus, recombination takes place close to the
wall and by this the concentration of atoms reduces.
Walls can have different surfaces. They can be made of metals or glass-like materials. This involves different reaction rates or time scales in the heterogeneous reaction
at the wall. Some values for the catalycity coefficient kw are printed below.
kw = 00
kw = 1 to 10m/sec
kw = 0.01 to O.lm/sec
kw = 0

{:}
{:}
{:}
{:}

only recombination
metal oxides
glass like material
no recombination

In case of a catalytic wall Figure 8 shows the concentration of oxygen atoms for
different values of the reaction rate kw at the wall.
Clearly the effect of a sink for atoms is visible with increasing catalycity. This
means that the recombination is the dominant reaction at the wall. Oxygen atoms
are forced to recombine and form oxygen molecules again.
In Figure 9 a graph of the downstream development of the wall concentration CAw
and the heat transfer to the wall qw is illustrated for different values of the catalycity
coefficient kw.
For small values of the catalytic wall reaction rate a steeper gradient of the wall
concentration is seen than for values up to infinity. In the latter case equilibrium
is achieved already very early. The other diagram shows heat transfer for zero and
infinitely high catalycity. The difference is in the double logarithmic scale 15% at
maximum.

Computational Methods in Hypersonic Aerodynamics 19

~---------.--------~

x=2m

4
y/8
3

~---------r----------4

r---~~~d---------~

Omfsec
5m1sec

50mfsec
~~--------~-i~-~loooo~~mf~se~c

0.05

c./c
-p,

0.10

Ae

Figure 8: Concentration profiles of oxygen atoms for different catalytic wall reactions (Ma

= 24)

0.10

Omfsec

---0.05

....... ........

....

..............

...........

----------

0.00
0.0

0.5

1.0

X/L

1.5

2.0

10 6

kw=

Omfsec

kw- 10000 mfsec

N/(m sec)

10 5 ~------------~------------~~~~~----_+------------~

10 4
10 1

X/L

Figure 9: Wall concentration and heat transfer at the wall along a flat plate (Ma

= 24)

20 Computational Methods in Hypersonic Aerodynamics

x=1m

x=1m

4
Pr = 1.0
Pr = 0.7

y/o

y/O

1
0

Figure 10:

(Ma

= 24)

15

TITe

30

0.05

e Ie
'A

0.10

Ae

Temperature and concentration profiles for variable Prandtl and Lewis numbers

Figure 10 presents the influence of the transport coefficients on the development


of temperature and concentration profiles. The coefficients are the Prandtl number Pr
and the Lewis numbers Le which involves the diffusivity of a gas. In literature these
non-dimensional numbers are often unity for reasons of simplification. In the figure
the differences are shown when more realistic values are preferred.
The temperature profiles are not very much influenced by the Lewis number but
by the Prandtl number. Higher Prandtl numbers causing higher temperatures make
the concentration of atoms increase. This is because the production density of atoms
increases exponentially with temperature. In Fig. 10 an increase by the factor of 2.5
is visible while keeping the Lewis number constant. The Lewis number when it is
enlarged causes a stronger diffusion of atoms to the wall near and outer regions of the
boundary layer. This effect is opposite to an increasing thermal conductivity involved
in the Prandtl number.
The effect of thermal diffusivity is regarded in Figure 11.
lt describes the transport of mass due to a temperature gradient. In the figure
calculations are presented with and without this effect. It turns out that in regions
of lower temperature, namely close to the wall or at the outer edge of the boundary
layer, the atomic component is transported due to the positive temperature gradient
to the location of the maximum temperature. Therefore rises the maximum amount
of atoms to about 14% and at the wall the increase is nearly 7%. This is computed
at a downstream position of 2 m of the flat plate.
The Dufour effect transporting energy due to a concentration gradient is of minor
order. The boundary layer results differ only by 0.2%.
In Figures 12 and 13 the concentration CAw and the heat transfer qw distributions
along the wall of a flat plate are shown for a sudden change in the catalytic surface

Computational Methods in Hypersonic Aerodynamics 21


material. The catalytic wall reaction rate of 50 m/s corresponds to a silver plated
surface. In the upper part of Fig. 12 where the non-catalytic portion of the wall is
located at the beginning of the flat plate the concentration of oxygen atoms decreases
rapidly when the surface material becomes catalytic.
This is because the catalytic part reacts like a sink for atoms. The dominant
reaction is the recombination. In the lower part of this figure with interchanged
catalytic wall regions the decrease of the number of atoms happens now more modestly
in the first part of the plate. In the second part, however, the concentration goes up
again.
The heat transfer plotted in Fig. 13 shows a rapid increase in the upper part of
the figure because of the sink behavior of a catalytic surface. The energy of the atoms
is transferred to the metallic portion of the surface. The value jumps three times
higher than the last value calculated at the non-catalytic part of the flat plate. It is
opposite in the lower part of the figure. Here the heat transfer is reduced by 1/3 of
the foregoing value.
These examples were given for an illustrative imagination of chemical reactions
in hypersonic boundary layers. The gas considered is a diatomic gas, namely oxygen.
The flow is supposed to be in thermal equilibrium but in chemical non-equilibrium.
A first impression is given by these boundary layer calculations along a flat plate
when it is exerted to very high, namely hypersonic velocities. In high temperature
regions these examples show the production of atoms which follow from dissociation
reactions. The destruction of atoms which comes from recombination reactions occurs
when the temperature decreases for any reason.
In the introductory part of this chapter it was stated that the temperature computed by means of simplified approaches was much too high. The temperature inside
the boundary layer flow allowing chemical reactions does not exceed the predicted
value of 11 000 K considerably. Depending on the precautions taken the surface temperature can even be reduced by an appropriate choice of the surface materials. The
inclusion of chemical reactions in the calculation of viscous flows demonstrates clearly
the influence on the temperature development.
The next section contains a nomenclature of the most frequently used definitions
necessary for reacting gas mixtures in high temperature flows.
6. DEFINITIONS FOR REACTING GAS MIXTURES IN HIGH TEMPERATURE
FLOWS
For the modeling of high temperature reacting gas mixture flows some obviously
necessary definitions of the flow properties have to be provided. A gas mixture is
composed of s species, which involve the same number of species mass fractions Cs.
Each one has a specific partial static pressure Ps. This shows in principle that many
new definitions have to be introduced in the modeling of the equations governing
chemically reacting flows. In this section the most frequently used definitions and
their interconnections are prepared.
In the low temperature range of aerodynamics a gas mixture like air is assumed
to be a perfect gas following the equation of state.

22 Computational Methods in Hypersonic Aerodynamics

x=2m
4

Vl6
3
2

Figure 11: Concentration profiles of oxygen atoms including the thermal diffusion effect (Ma

0.50

cAwIe Awe
0.25

'W= 0 mlse~
t-

--

f---

I
I
I
I
I
I
I
I

K"v= 50 m/s

= 24)

I
I

0.00
0.0

0.1

0.2

0.50

'W= SOmis IIC

cAw Ie Awe
0.25

0.00

I
I
I
I
I
I
I
I

0.4

X/L

0.5

0.6

'V,,= Om/se

I
I
I
I
I
I

1"'0.0

0.3

0.1

0.2

0.3

0.4

X/L

0.5

0.6

Figure 12: Wall concentration of oxygen atoms along a flat plate with variable catalycity (Ma

= 24)

Computational Methods in Hypersonic Aerodynamics 23

,,

4.0

kw= Om/se<

'\At 10-6
N/(m sec)
2.0

"'--

~w= som/sep

1\

r--

,,

0.00

0.1

0.0
0.50

~ 10-6

N/(m sec)
0.25

0.2

~w=

0.3

0.4

X/L

0.5

0.6

,,

SOm/se:;

kw= Om/se

,
,

"-----

---

,,
,,
,

L
,

0.00

0.0

0.1

0.2

0.3

0.4

X/L

0.5

Figure 13: Heat transfer at the wall along a flat plate with variable catalycity (Ma

p=pRT

0.6

= 24)

{:} equation of state of air

where R is the specific gas constant of air having the value:


m2

R = 287,1 s21<

{:} specific gas constant of air

This equation of state sets the pressure p in relation to the density p and the
temperature T.
The definition of a perfect gas contains the assumption that no intermolecular
forces are acting. These forces are described for example by the LENNARD-JONES
potential CPm which is a function of the distance r from the gravitating body.

where d is the characteristic diameter of the molecule and t: is the characteristic energy
of interaction between molecules. This potential is related to the intermolecular forces
by differentiating it with respect to the distance r.

F: __ dCPm
m dr

24 Computational Methods in Hypersonic Aerodynamics


Since in gases the distance between the molecules is large especially when the
temperature is high, these forces are of negligible order. The influence of the attractive
force becomes negligible about 10 molecular diameters away from the molecule. Dense
fluids in very low temperature regions develop considerable intermolecular forces. In
cases like these where the gas does not behave in a perfect way anymore, the equation
of state by VAN DER WAALS must be applied. The gas becomes a real gas. In
contrast to some literature, real gas effects are related to intermolecular forces and
not to high temperature effects, see also ANDERSON [3].
Coming back to high temperature effects where molecular distances are much
larger, other effects than these attraction force effects must be considered about the
equation of state. Now, the air will be seen as a combination of different perfect gases
which have their specific properties. So is the specific gas constant R. different for
each species s in the perfect gas mixture. The specific gas constants R. then must be
related to the universal gas constant R and the molecular weight M. in the following
way.
R
R.= - -

M.

where M. is defined also as the mass of a species s per mole of this species which is
measured in kg/kg-mole. The unity kg-mole is one single expression which cannot be
split into parts. The universal gas constant has the value:
R = 8.3144 . 103

m 2 kg
I {:} universal gas constant
g - mo e

2 I< k

The AVOGADRO's number N A is the number of particles per mole. It connects


the universal gas constant R with the BOLTZMANN number k also called the gas
constant per particle.
{:} Boltzmann number

These numbers are constant and have the following values:


2 particles
N A = 6.02 . 10 6 k
I {:} A vogadro number
g- mo e

k = 1.38 . 1O-23::~Z.

{:} Boltzmann number

Now, in case of a gas mixture consisting of several species the equation of state
IS:

p=

E p.R.T = E.~, RT

{:} equation of state of gas mixture

where P. and Ms are the species density and the molecular weight. The latter also is
referred to as the mass per mole of species s. Finally, from this state equation it turns

Computational Methods in Hypersonic Aerodynamics 25


out that the pressure of the mixture is composed of the sum on the partial pressures
p. of the single species.

p=

L:P.

The molecular weight of the gas composition M can be obtained by simply reintroducing the perfect gas equations for the mixture as well as for the species.

R=

L: P. R.
P

The definition given above of the universal gas constant


tion finally yields:
1

M=-L.. -it

n introduced into this equa-

{:} molecular weight of mixture

Since the gas in chemically reacting flows has to be considered as a composition


of several gases, some definitions for their masses, volumes and densities and other
properties must be provided. These definitions and their interconnections are listed
below including the dimensions used for non-dimensionalization purposes.

{:} mass of mixture in kg

M.

{:} mass of species s in kg

{:} number of particles of mixture

N.

{:} number of particles of species s

{:} volume of mixture in m 3


{:} volume of species s in m 3

v -M
- y.
V

-.Y.t..

M.

{:} specific volume of mixture in m 3 / kg


{:} specific volume of species s in m 3 / kg
{:} density of mixture in kgJm 3
{:} density of species sin kg/m 3
{:} molecular weight of mixture in kg J kg-mole

{:} molecular weight of species s in kg J kg-mole


.r_M
M

JV

Ar

JV. -

&

M.

V=~

{:} number of moles of mixture in kg-mole


{:} number of moles of species s in kg-mole
{:} molar volume of mixture in m 3 J kg-mole

26 Computational Methods in Hypersonic Aerodynamics

VS

-x...
N.

molar volume of species s in m 3 / kg-mole

There are different ways of describing the concentrations of the various species in
a multi-component gas mixture. Obviously, they differ from reference to reference.
The most common ones are presented only.

Ms
Ps= V

mass concentration of species s in kglm3


(mass of species s per unit volume of mixture)

cs =~
Ms

molar concentration of species s in kg-molelm3


(moles of species s per molecular weight of species s)

The density of a species s is designed by Ps and Ms is the mass, while Ms is the


molecular weight. V indicates the unit volume of the gas mixture. Dividing these
definitions by the mass concentration P and the molar concentration C of the gas
mixture represents the mass and the molar fraction of the species s. These variables
are non-dimensional and frequently applied.

Ps

cs = -

ps

=-

mass fraction of species s (mass of species s


per unit mass of mixture)

_ Cs

Xs -

Ps
p

--

mole fraction of species s (moles of species s


per total moles of mixture)

Due to the given definitions they can be rewritten in terms of the partial pressure
divided by the pressure of the gas mixture.
A frequently appearing ratio is the so-called number density which expresses the
number of particles of the species s in a unit volume of the gas composition.

Ns

ns= V

number density in particles of species slm3


(number of particles of species s per unit volume of mixture)

Another conclusion of these definitions is that:

Computational Methods in Hypersonic Aerodynamics 27


{:} mass density of gas mixture

{:} molar density of gas mixture


a

{:} unity for gas mixture


s

Two expressions are often used to define a mean velocity of the gas composition
usually appearing in the momentum or global continuity equations. One supports a
mass averaged velocity Ui and the other a molar averaged velocity Ui. The index i
stands for the three velocity components in the three spatial directions.
{:} mass average velocity

{:} molar average velocity

The species velocity is shown by Usi . This velocity is meant to be the sum on
all the velocities of the molecules of species s inside of a very small volume in the gas
composition, divided by the number of these molecules. Thus, the species velocity
and the average velocity are measured with respect to a stationary coordinate system.
The mass averaged velocity is the most frequently applied in fluid mechanical
literature. The molar way of averaging will be omitted here, although all the further
relations can also be performed in this way.
The difference between the species and mean flow velocities is called the diffusion
species velocity indicated by Vai. This means that each species moves in the gas mixture
flow with its individual speed. The diffusion velocity is measured with respect to the
mass averaged velocity. The index i again shows that there are three spatial directions
involved.
Vai

= Usi - Ui

{:} mass averaged diffusion velocity of species s

The diffusion velocity represents a measure of the relative motion between the
gas mixture in total and its single species. Multiplying this mass averaged diffusion
velocity by the species density Ps defines the mass flux of the species s. It is the
diffusion flux of this individual species in the gas composition of different species.

jsi

= paVsi = Pa(Uai - Ui) {:} species mass flux with mass averaged velocity

This species mass flux jai is a vector quantity. It is the mass of the species s
passing through a unit area per unit time, or a partial mass flux per unit volume Pa
with the corresponding velocity Vai.

28 Computational Methods in Hypersonic Aerodynamics


A special property of the mass fluxes is that the sum on all the species mass fluxes
obviously has to be zero.

Ljs;

= 0

This section on the definitions of properties in reacting gas mixture flows closes this
chapter on some introductory remarks to hypersonic aerodynamics.

REFERENCES
1. U.S. Standard Atmosphere 1962. NASA, USAF, USWB, US Government, Printing Office, Washington DC, 1962.
2. Anderson, J.D., Jr. Modern Compressible Flow, McGraw-Hill, Second Edition,
1990.
3. Anderson, J.D., Jr. Hypersonic and High Temperature Gas Dynamics, McGrawHill, First Edition, 1989.
4. Anderson, J.D., Jr. Introduction to Flight, McGraw-Hill, Third Edition, 1989.
5. Hankey, W.L. Some Design Aspects of Hypersonic Vehicles, AGARD-LS 42,
Aerodynamic Problems of Hypersonic Vehicles, 1972.
6. Peters, N. Losung der Grenzschichtgleichungen fur chemisch reagierende Gase
mit einem Mehrstellenverfahren, DLR-FB 72-58, 1972.

Chapter 2:
Computational Methods for Viscous Hypersonic
Flows
D. Hanel
Department of Combustion and Gasdynamics,
Universitiit-GH-Duisburg, D-4100 Duisburg, Germany
ABSTRACT
This paper is concerned with computational methods for the solution of the NavierStokes equations for hypersonic flows. The governing equations and relations for
equilibrium flow are described in the first part. Their numerical formulation is outlined
including a brief consideration of the conservative discretization, the evolution of the
numerical fluxes and of typical methods of solution. As a further topic the influence
of the numerical damping on the solution is discussed demonstrated for steady-state
solutions of the Navier-Stokes equations. The special properties of hypersonic, viscous
flows and the resulting numerical problems are considered finally.
INTRODUCTION
The computational fluid dynamics is an essential tool for the design of spacecrafts
and hypersonic airplanes. The physical and numerical problems to be solved cover
a wide range of the fluid mechanics, which are ranging from rarefied gasdynamics,
chemical non-equilibrium to problems of non-reactive, continuum flow. Rarefaction
and chemical effects are of great importance, in particular for reentry vehicles. Large
portions of the flight path are governed by continuum flow, in particular for hypersonic
planes in the Mach number range of five to ten, but also for reentry bodies in the
lower atmosphere.
The present paper is concerned with the computation of viscous, hypersonic flow in
the continuum flow regime. The viscous flow is of great importance for the prediction
of the flight properties, in particular for the prediction of aerodynamical forces and
heat flux rates. The vehicles considered are usually blunt bodies or planes with
a low aspdct ratio, thus the flow is three-dimensional and strong viscous/inviscid
interaCtions can appear. The typical flow around blunt bodies is characterized by
strong shock waves and expansions, embedded subsonic regions, and shock-boundary

30 Computational Methods in Hypersonic Aerodynamics


layer interactions with separation. To attack such flow problems the Navier-Stokes
equations have to be solved. The numerical methods required for the solution of such
a problems must be powerful, should be sufficiently accurate in viscous flow regions
and should have the properties of high resolution schemes in the nearly inviscid flow
portion.
From the numerical point of view essential difficulties can arise from the structure
of the Navier-Stokes equations, which involves different characteristic scale lengths
to be resolved numerically. This fact requires much finer computational meshes than
needed for the corresponding inviscid problem, and with that more computer storage and computational work is required. Besides a necessary increase of computer
capacity, the efficiency of numerical methods has to be improved.
Another aim in the development of Navier-Stokes solvers is the improvement of
the accuracy of viscous solutions, in particular for flows at high Reynolds numbers. In
this case there is a very sensible balance between inertia and viscous terms. Physically
the inertia terms have no dissipative contribution, but their numerical approximation
generates a certain amount of numerical dissipation superposing the physical dissipation. Then the accuracy of the solution can be impaired, in particular in viscous layers
where strong gradients are present. These problems, important in every numerical
solution will be discussed in one section.
The special numerical problems arising in hypersonic, viscous flows will be discussed in a further section, and by means of different approaches as used in present
applications.
MATHEMATICAL FORMULATION OF THE FLOW PROBLEMS
Different flow ranges can be observed during the reentry of a spacecraft. At high
altitudes low density effects become important. The flow is essentially a free molecular
flow due to the increased mean free path between the molecules. At lower altitudes
the flow becomes more and more collision-dominated, and approaches continuum flow.
From the gaskinetical point of view, the different flow ranges can be characterized by the Knudsen number K n, which is defined as the ratio of the mean free path
between the collision of two molecules, and of a characteristic macroscopic (body)
length. According to the value of the Knudsen number the flows can range from the
nearly collision-free molecular flow (K n ~ 1) to the collision-dominated continuum
flow for K n ~ 1. The governing equation for all these flow ranges is the Boltzmann
equation, an integro-partial differential equation for the molecular distribution function. However the complexity of this equation is high, and numerical solutions become
very costly (e.g. the Monte Carlo methods).
For most situations encountered at ordinary densities in gas dynamics the Knudsen number is very small, the corresponding flows are continuum or near-continuum
flows. Therefore the solution of the Boltzmann equation can be avoided by expanding
the Boltzmann equation with ~espect to small Knudsen numbers (Chapman-Enskog
expansion). By that a hierarihyof flow equations can be derived with increasing order
of Kn, which gives in increasing order the Euler equations, the Navier-Stokes equa-

Computational Methods in Hypersonic Aerodynamics 31


tions, and the Burnett equations (e.g. see Vincenti, Kruger [1]). Equivalently with
increasing Knudsen number, the hierarchy describes the increasing deviation from the
state of thermodynamical (translational) equilibrium. In this sense the Euler equations of inviscid flow can be considered as the conservation laws for thermodynamical
equilibrium, whereas the Navier-Stokes equations describe small deviations from that.
Both the Euler and Navier-Stokes equations are the most important tools for predicting technical relevant flow problems. The Burnett equations, although higher order
equations have shown only small, if any improvement over the Navier-Stokes equations.
Theoretical and experimental investigations have confirmed that the Navier-Stokes
equations describe sufficiently well the flow even in the near-continuum flow range (not
to far from continuum). Furthermore these equations, although derived for nearly
thermodynamical equilibrium, are equally valid for flows with vibrational or chemical
non-equilibrium, however, provided the thermodynamical variables are given in their
extended definition appropriate to the non-equilibrium situation.
Conservation laws
In the present paper continuum flow at thermodynamical and chemical equilibrium
is considered. The flow is described by the conservation equations for mass, momentum, and energy. These conservation quantities are expressed in a vector Q with the
components of specific quantities per volume, which are the specific mass (density
p), the momentum pv ,and the energy pE. The conservation laws are formulated as
the rate of change of the conservative variables Q in a control volume T, which is
balanced by the effect of the generalized fluxes H acting normally on the surface A.
The conservation laws in the integral form read:

(1)
A differential form (divergence form) can be obtained with the integral theorem of
Gauss:

(2)
Herein Q = (p, pv, pEf is the vector of the conservative variables. The generalized
flux H can be split into a vector for inviscid flow Hinv and a vector H vi8C describing
the contribution viscosity and heat conduction on the flow. The terms read:

H-inv

= (pv, pv- v- + p, pv-H)T


t

and

HVisc

(0, a, a . v+ iff

(3)

where is the stress tensor, and ifis the heat flux vector.
For the sake of simplicity in the following discussion the conservation equations
and their approximations are written for a two-dimensional Cartesian coordinate system (x,y,t).
Full Navier-Stokes Equations The most complete description of continuum flow is
given by the N avier-Stokes equations. This system of equations is formed by the laws
of conservation of mass, momentum and energy for viscous, heat conducting fluids.

32 Computational Methods in Hypersonic Aerodynamics

In a Cartesian frame the flux vectors are split up in Cartesian components, i.e.
H- inv = (F, G) T and Hvi.
= (5, R)T
. The'mtegral form now reads

(4)
and the corresponding divergence form gives

(5)
Herein is

where 84 = qx + UTxx + VTxy and r4 = qy + UTxy + VTyy . With the Stokes assumption
/1v = -2/3/1 the stress terms and the components of the heat flux vector are:

aT
ay

qy =A-

The Reynolds averaged Navier-Stokes equations for turbulent flows show the same
structure and therefore are treated in the same way.
Thin Layer Approximation The Thin Layer Approximation of the Navier-Stokes equations is a widely used approximation for the computation of viscous flows at high
Reynolds numbers. Similar as in the boundary layer theory all the viscous terms with
stream-wise derivatives are neglected for the Thin Layer Approximation. However, in
contrast to the boundary layer theory, the Thin Layer Approximation retains all the
terms of the Euler equations, and (-,e time derivatives as well. Consequently, the Thin
Layer Approximation preserves all the properties of the inviscid flow, in particular the
information transport along characteristics and discontinuous solutions.
In 2-D Cartesian coordinates, assuming the x-coordinate as the nearly streamwise
direction, the Thin Layer Approximation would read:

(6)
where Q, F, and G have the meaning as in the full equations, but the viscous term
R contains only those stress terms which have derivatives in normal (y-) direction.
There is no rigorous theory for the derivation of this approximation, but its range
of validity can be considered approximately the same as that of the higher order
boundary layer theory. It means that local flow separation and small normal pressure
gradients in viscous layers are covered by the approximation.
A further motivation for using this approximation is given by the fact that very
different scale lengths exist in a boundary layer, and therefore the step sizes in streamwise direction are much larger than in normal direction. By this, even when the full

Computational Methods in Hypersonic Aerodynamics 33


equations are used, the streamwise derivatives of the viscous terms cannot be resolved
sufficiently well. In Principle, also most of the turbulence models in applications are
suited only for boundary layer-like flows, and therefore are satisfied by the Thin Layer
approximation.
An important requirement for the application of the Thin Layer approximation
is the use of streamline- (surface)- orientated, orthogonal meshes in viscous layers to
resolve completely the remaining main stress terms normal to the surface.
The numerical methods of the solution for the full Navier-Stokes equations, and
the Thin Layer approximation are nearly identical.
Parabolized Navier-Stokes Equations The Parabolized Navier-Stokes equations correspond to the Thin Layer approximation, but usually in their steady-state form. For
the example of 2-D flow the parabolized equations would read:

Fx

+ (G -

R)y = 0

(7)

where the viscous terms R contain only terms with derivatives normal to the main
flow, F, and G are the complete Euler fluxes.
Efficient space marching methods are the motivation for using this approximation.
Space marching is well suited for stationary, supersonic flow, where all information
is transported downstream within the Mach cone. But in the subsonic regions this
assumption fails, since the stationary equations become elliptic and upstream influence
occurs in main flow direction. Therefore the term "parabolic" is somewhat misleading.
To preserve the "parabolic" behaviour, space marching in main flow direction must be
enforced numerically by one-sided differences for all derivatives in main flow direction.
This numerical manipulation can be justified by the "parabolic" nature of attached
boundary layers. However neglecting the upstream influence in subsonic regions,
numerical instabilities can arise, which must be suppressed. This can be done with
different strategies.
In common marching procedures for external flows either the pressure gradient
normal to the wall is assumed to be zero across the subsonic layer, Schiff and Steger
[2], or the contribution of the streamwise pressure gradient is decreased in the subsonic
layer as a function of the Mach number based on a stability analysis, as proposed by
Vigneron [3].
Space marching methods can also be constructed for the time-dependent Thin
Layer equations, Eq.(6), using an time marching technique. The solution converges in
time in an iteration-like manner for each cross flow plane separately. A usual upwind
scheme can be used, where all of the variables needed downstream of the actual cross
section are extrapolated from the upstream sections without further assumption for
the pressure, as reported by Menne [4] for computations of external hypersonic flows.
An advantage of this strategy is that the same time-marching code can be used in
"parabolic" and "elliptic" regimes.
Since only a single space marching sweep is employed in fully parabolic approximation procedures, the computations become very efficient with respect to the computation time.

34 Computational Methods in Hypersonic Aerodynamics


Like the boundary layer approximation, the parabolic assumption is not able to
deal with flow separation in streamwise direction. In normal direction however, cross
flow separation and strong secondary flow can be predicted by this assumption. A
sensitivity study in parabolized Navier-Stokes solutions of external supersonic flow was
carried out e.g. in [5] with the aim to extend this approximation to more demanding
flows at high angle of attack and to improve the accuracy and stability.
The application of parabolic procedures requires initial conditions in two or three
cross sections. These conditions have to be taken from other solutions or from experiments.
Reduced Navier-Stokes Equations The Reduced Navier-Stokes Equations (RNS) correspond to the Parabolized Navier-Stokes equations, but the approximation of the
streamwise pressure gradient term is modified to account for the upstream propagation of pressure waves within subsonic zones. In this way the pressure is treated as
elliptic and stored in the whole domain. The remaining variables are treated as in
the fully parabolic approximation and stored only on some cross section as in space
marching procedures. Due to the elliptic treatment of the pressure the approximation
enables the calculations of weak streamwise separation and is applied to external and
internal flow problems.
The RNS solution has shown to be somewhat more expensive than the PNS solution, but much cheaper than the full Navier-Stokes solution as demonstrated e.g. by
Power, Barber [6] for external flow over a compression ramp.
Conical approximation The conical approximation is often used to generate an initial
condition for parabolic space marching methods. The conical approximation presumes
similarity of the solution in cross sections along the axis of a cone (3-D) or of a wedge
(2-D). Similarity means, the solution remains constant along rays through the origin
of the cone. Transforming the conservation laws in conical coordinates and using this
assumption, the spatial dimension of the equations is reduced by one order, i.e. from
3-d to 2-D and from 2-D to I-D.
The conical approach is correct for inviscid flow over a sharp cone with attached
shock. For viscous flows this approach is not valid in a general way, but is often used
as a local approximation to compute an initial condition for the first cross section in
a space marching method.
Euler equations The Euler equations, describing the inviscid flow, are an important
approximation of the Navier-Stokes equations. They contain the essential mathematical difficulties and therefore determine the properties of Navier-Stokes solutions.
Nearly all of the Navier-Stokes solvers for high Reynolds numbers are based on Euler
solvers extended by the viscous terms.
The 2-D Euler equations read in their conservative integral and divergence form:

+ Fx + G y = 0
Qtdr + Fdy Gdx =
Qt

(8)
0

(9)

Computational Methods in Hypersonic Aerodynamics 35


The Euler equations form a nonlinear, hyperbolic system of equations with real eigenvalues AI. As a consequence of the nonlinearity, the equations show two different types
of solutions, discontinuous (weak), and continuous, smooth solutions, as well.
The continuous solution can be expressed by the characteristic solution of the
Euler equations. The characteristic form of these equations is obtained by a diagonalization of Jacobian of the Euler fluxes, e.g. A = ~~ with the corresponding eigenvector
matrix T . With the diagonal matrix A = diag()../) = T-1 AT the characteristic form
reads in I-D:

(10)
The characteristic variables Ware defined by dW = T- 1 dQ and the diagonal matrix
is given by the eigenvalues, A = diag( u + a, u, U - a) . The characteristic form is
the basis of the method of characteristics, but also the basis for constructing upwind
shock capturing schemes.
The weak solution, describing the jump conditions over a discontinuity (e.g. a
shock wave or a slip line), can only be derived from the conservative integral form
of the Euler equations. Therefore, if embedded discontinuities are considered, only
the conservative form guarantees the correct jump conditions. For a discontinuity C
moving with the velocity c, the application of the integral conservation laws results in
the jump conditions which read in the general form with the definition [Jl = 12 - 11:

L[H - Qc] iidA

= 0

(11)

By means of this jump condition the Rankine-Hugoniot relations can be derived, and
computationally shock-fitting procedures can be constructed.
Thermal and calorical relations
The solution of the conservation laws requires additional closure relations to express
the thermal and calorical state, and the transport quantities in the flux Ii as function
of the conservative variables Q.
Different situations have to be considered for the formulation of thermal and
calorical closure relations. The situations are equilibrium, frozen and nonequilibrium
flows. For the present consideration the gas is assumed to be in thermodynamical and
chemical equilibrium.
Assuming equilibrium flow, the equations of state can be expressed as algebraic
closure relations for the thermodynamical and calorical state as function of conservative flow variables Q. The basic input quantities for these relations usually are the
density p, and the internal energy c:, which can be calculated from the total energy

pE = p(c: + ij2/2).
The caloric equation of state expresses the internal energy c: with two thermodynamical variables, e.g. with p and T.
c: = c:(p, T)

36 Computational Methods in Hypersonic Aerodynamics


For real gases, i.e. a mixture of gases at equilibrium, the caloric equation of gases
is the sum of all energy contributions of the species. The single contributions can
be calculated with methods of the statistical thermodynamics and with data from
measurements (see Vincenti, Kruger [1]).
For a thermally perfect gas (e.g. air T ::; 2000I<) the caloric equation reduces to
c = c(T) = J cvdT, and with the additional assumption of calorically perfect gas (e.g.
air t ::; 600) it becomes c = cvT.
The thermal equation of state defines the pressure as function of two variables, e.g.
of the internal energy c, and of the density p. Assuming that each species of a mixture
behaves like perfect gas, the pressure becomes the sum of the partial pressures Pi of
each species:
R
(12)
P = LPi = LPi- T = p(p,c)
i

Mi

For a thermally perfect gas, e.g. air without dissociation, it yields P = pRT. In
addition for a calorically perfect gas it is P = h - 1) pc.
With the equations of state known, and with the thermodynamical laws all the
other quantities can be derived. In general the equations of state for a real, nonperfect gas cannot be formulated in a closed form. The calculation of the state has
to be carried out e.g. by an evaluation of the thermodynamical partition functions,
from which Mollier diagrams can be constructed.
Computationally it results in a system of nonlinear, algebraic equations, which has
to be solved for each state (grid) point. For computational purposes the precalculated
variables of state can be subdivided in regions in the p, c plane and expressed by curve
fits in those regions (e.g. for air by Tannehill [7]). For improving the vectorization
the equation of state can also be used in form of interpolation tables with equidistant
or stretched stepsizes in the p, c plane (e.g. Vinokur [8]).
Transport Quantities
The computation of viscous, heat conducting flow requires additional relations for the
transport coefficients of momentum, (viscosity), of energy (thermal conductivity), and
in mixtures of gases for the diffusion of species masses (diffusivity). For the situation
of chemical equilibrium, as considered here, the diffusion of species is usually assumed
to be negligible.
Laminar flow The transport quantities in laminar flow are functions of the molecular
properties, depending on the local thermodynamical state. In general the values and
the dependence are well known.
At moderate densities the coefficients of viscosity and heat conduction of a single
perfect gas are functions of the temperature only. Then simple models can be used
for the calculation of the viscosity of a single component gas, as e.g. the Sutherland
formula or potential laws. The corresponding thermal conductivity can be calculated
from the viscosity using the Eucken relation (Vincenti, Kruger [1]).
For a mixture of gas at equilibrium, where each component is assumed to be a
perfect gas, the composite viscosity can be determined from the semi-empirical mixing

Computational Methods in Hypersonic Aerodynamics 37


law of Wilke, which takes into account the species viscosity and the corresponding
concentration. For the thermal conductivity of a gas mixture a similar law can be
defined (for details see e.g. Bird, Stewart, Lightfoot [9]). With such mixing laws the
transport coefficients become functions of two variables, e.g. J.l = J.l(p, c). Therefore
the calculation of transport coefficients has to be combined with the calculation of the
thermal and calorical equations of state. Computationally the transport coefficients
can be updated similar to the equations of state in form of interpolation tables or
curve fits (e.g. for air by Srinivasan [10]).
Turbulent flow The problem of closures in turbulent flows is much more complicated
than for laminar flow. Generally all turbulent flow simulations suffer from a lack of
correct physical modelling. The complexity of describing turbulent flows is still larger
for hypersonic flows due to the effects of compressibility, threedimensionality, and
nonequilibrium mixtures.
The common attempt to solve the Navier-Stokes equations is based on timeaveraging of the variables, which results in the Reynolds-averaged Navier-Stokes equations. Additional closure relations have to be found between the averaged fluctuating
quantities (Reynolds stresses), and the mean values of the conservative variables.
A usual formulation of the turbulent stress terms is the eddy viscosity concept,
which models these terms similar to the laminar case as an eddy viscosity (turbulent
viscosity) multiplied by the velocity gradient. Defining an effective viscosity, J.lef f =
J.llam + J.lturb, the Reynolds-averaged Navier-Stokes equations can be written in the
same form as for the laminar flow. Thus the mathematical structure of equations is
retained, and the numerical methods of solution do not differ essentially for laminar
and turbulent flow, beside of the strongly nonlinear transport coefficients and the
additional closure relations.
Closure relations can be achieved by additional transport equations, like the k - c
model, or by algebraic relations, like the mixing length assumption. All of these
assumptions need parameters to be adapted empirically to the special flow problem.
The most of these models are suited for boundary layer-like flows with small separation
zones only (which motivates the use of the Thin Layer approximation).
In Navier-Stokes computations an additional difficulty arises from the evaluation
of the boundary layer thickness, necessary for the scaling of the models. Therefore
special turbulence models were derived for the requirements of Navier-Stokes solutions.
At present two widely used models are the two-layer algebraic model by Baldwin and
Lomax [11] and the k - c model. Details can be found in the related literature, a
review of turbulence closures in Navier-Stokes solutions is given e.g. in [12].
An additional, and essential problem arises from the prediction of the transition
from laminar to turbulent flows. This problem is much more uncovered than the
turbulent closures, and requires further investigations in particular for compressible
high-speed flow.
Boundary conditions
The boundary conditions at the integration domain define the specific problem to be
considered. Therefore they are an essential part of computation. In most physical

38 Computational Methods in Hypersonic Aerodynamics


situations they can be divided into wall conditions, in- and outflow conditions, and
periodic conditions. In this paper continuum flow at thermodynamical equilibrium is
considered. Thus the boundary value problems correspond essentially to that known
for lower Mach numbers.
Periodic boundary conditions utilize the periodicity of flow field. Their formulation
is usually straightforward.
The boundary conditions at the wall for continuum flow are defined by the conditions
of vanishing normal velocity, by the no-slip condition (vanishing tangential velocity
vt}, and by the thermal conditions:
aT = 0
or
vn = 0,
(13)
Vt = 0,

on

In the transition regime (not too far from the continuum) the flow is still sufficiently
represented by Navier-Stokes equations. The influence of rarefaction effects can be
restricted to the boundary conditions at the wall by using the so-called slip conditions (or incomplete accommodation of momentum and energy). Then the boundary
conditions for the temperature and the tangential velocity form a system of coupled
equations for the value, and for the gradient at the wall. They can be written in a
generalized form:

ali
AV+B-=C

an

with

(14)

The coefficients in A, B, and C are given by the gaskinetical theory. An application


of these slip conditions for viscous, near-continuum flow is given e.g. by Gokcen, Mac
Cormack, Chapman [13]. In this paper it is shown that the Navier-Stokes equations
are able to describe the transition flow by using appropriate slip boundary conditions.
As an example from [13], Fig. 1 shows the computed skin friction drag versus Reynolds
number on a flat plate using different wall conditions.
The inflow and outflow boundary conditions are less unique, since they depend stronglyon the flow problem considered. The number of the boundary conditions (conditions from outside) and of the compatibility conditions (conditions from the interior
integration domain) can be derived from the differential problem. For the complete
Navier-Stokes equation such conditions are discussed e.g. in [14], [15]. For the thin
layer approximation additional boundary conditions due to viscous terms are not
required.
In many cases the correct conditions cannot be satisfied, therefore the assumptions
of nearly inviscid flow (characteristic updating of the conditions) and of the boundary
layer concept (parabolic behaviour in space) are often used.
The inflow boundary conditions for hypersonic flow are usually well defined by the incoming supersonic flow forming the bow shock in front of a body. If the bow shock is
captured the boundary conditions are represented by the supersonic inflow conditions.
Since in Navier-Stokes computations the viscous shock structure is not resolved, the
shock can be treated as a discontinuity and fitted as in inviscid flow.

Computational Methods in Hypersonic Aerodynamics 39


The outflow boundary conditions are more cumbersome. Usually the outflow boundary cuts an unknown flow field, and reasonable conditions have to be deduced from
the physical problem. If the inviscid flow is supersonic and if the viscous portions at
the outflow boundary are boundary layer-like, then the extrapolation of the boundary
values from the interior is a reasonable condition. In other cases additional conditions
from outside (e.g. the pressure) have to be prescribed.
SPATIAL DISCRETIZATION
The complete Navier-Stokes equations form a system of quasi-linear partial differential equations of parabolic-hyperbolic type in the time-space plane, and of elliptichyperbolic type in the space (steady-state). These different types demonstrate the
complexity of the solutions.
However the main problems in numerical solutions arise essentially from the many
disparate length scales and time scales, which are characteristic for the different physical phenomena. Typical length scales are for example the body length L, the boundary
layer thickness Db rv L /.../Re-, the viscous thickness of a shock wave D. rv L / Re, and the
scaling of turbulent eddies Ot rv LVRe 2 These scalings differ by orders of magnitude
in ordinary flows. Since not all of them can be resolved, the unresolved scalings must
be modeled, either by physical closures (e.g. turbulence) or by numerical means (e.g.
shock capturing schemes), otherwise the accuracy and convergence of a numerical
solution is impaired.
For most external and internal flows of technical interest the two most important
flow regions are the viscous layers with Db rv L /.../Re-, and the nearly inviscid flow with
scaling of order of L.
The viscous layers (boundary layers, wakes) are characterized by a small extension
Db normal to the main flow. Within these layers the Euler and viscous terms are nearly
balanced. Thus the solution becomes continuous, but shows large gradients and a
strong curvature of the variables. This fact can lead to severe problems of accuracy
of the numerical method if these layers are not resolved sufficiently well.
In the nearly inviscid regions the Euler terms are dominating and the properties
of the solution correspond essentially to that of the hyperbolic Euler equations. Thus
continuous and discontinuous solutions can appear as well.
Although viscous solutions are continuous in principle, discontinuous solutions
appear numerically, since in general the viscous structure of a shock wave with the
much smaller scale length is not resolved. Therefore the capturing of the discontinuous
solutions requires the use of the conservative form of the Navier-Stokes equations and
the use a Euler solver with good shock capturing properties.
Thus the Navier-Stokes solution includes all the numerical problems of a Euler
solution, and much care has to be taken to avoid undesired interactions between the
numerical dissipation of the Euler solver, and the physical dissipation from the viscous
terms in the N avier-Stokes equations.
The numerical solution of the conservation equations is based on a discrete approximation of the equations and their initial and boundary conditions as well. The

40 Computational Methods in Hypersonic Aerodynamics


aim of the numerical solution is to achieve an approximative solution, which converges
with decreasing step sizes to the analytical solution of the differential problem. The
proof of the convergence for the nonlinear initial-boundary value problems is rather
difficult. Therefore in most cases convergence is assumed, if the equivalence statement by P. Lax [16] is satisfied for a linearized version of the scheme. This proof,
valid for initial value problems, requires numerical stability and consistency as necessary and sufficient conditions for convergence. Both requirements are fundamental
for the development of converging methods.
The development of methods of solution for conservation equations can be divided
into different steps. These steps are the grid generation and arrangement, the conservative formulation in the discrete space, the spatial discretization of the fluxes and
boundary conditions, the time discretization, and the solution of the resulting system
of algebraic equations.
Conservative discretization
The numerical solution requires the preservation of the conservation properties in the
discrete space. The conservation equations are applied to a finite control volume (element), which is defined around or between the grid points. The numerical techniques
for a conservative spatial discretization differ essentially in the arrangements of the
control volume and in the discretization approach.
Arrangements of the finite control volumes in a given grid can be defined in
different ways.
Commonly used arrangements are the "node-centered", the "cell-centered", and the
"cell-vertex" arrangements, which are sketched in Fig. 2. For node-centered and cellvertex schemes the variables and the geometry are defined on the same grid point,
for cell-centered schemes the variables are defined in the center of the cell. These
arrangements of control volumes can be extended straightforward to triangulated,
unstructured grids as e.g. in [17].
In a Cartesian grid these arrangements show nearly the same accuracy, but for
a curvilinear grid different truncation errors can result. For example using a central
scheme it could be shown that for a skewed grid the error from the "cell-centered"
arrangement is larger than that for the "cell-vertex" arrangement [18]. Also the nodecentered scheme shows a better truncation error for highly skewed meshes, [19].
Near a boundary the different grid arrangements require a different numerical
treatment of the boundary conditions, which can be important for the spatial accuracy,
in particular near geometrical discontinuities like the trailing edge [19].
Although the truncation error analysis has shown different behaviour for the different arrangements of control volumes, sufficient accurate results were achieved with
all of them. The influence of the grid arrangement in combination with the numerical
discretization is very complex, therefore no general recommendation can be given for
the choice of any arrangement.
Conservative discretization of the governing equations can be achieved by using
the integral form, and divergence form as well. In the literature different discretization
techniques are applied.

Computational Methods in Hypersonic Aerodynamics 41


The "Finite-Volume" approach uses the integral form, Eq. (1). The surface
integral is approximated by the sum over all faces of the control volume. In a simplified
form the discrete form reads in 2-D:
~Q

~Vol

ut

+ L:(F~y -

G~x)

(15)

k=l

Herein Q is the volume-averaged value, (F~y - G~x) are the normal fluxes summed
up over the single cell interfaces. The advantage of the "Finite-Volume" approach
is the direct application to the physical (x,y) space, and the easy interpretation in
curvilinear meshes. This approach can be used for structured and unstructured meshes
as well.
The "Finite-Difference" approach, based on the divergence form, Eq. (2), is
slightly more expensive in its evolution, since first the differential equations have
to be transformed to the curvilinear coordinate system and afterwards the equations
are discretized in the transformed plane. The advantage of this approach is given for
the more complex terms like the viscous terms in the full Navier-Stokes equations.
For the 2-D example the transformed equations in a curvilinear coordinate system
(~,1J) read:

(16)
with Q
QJ and F
FY7J - Gx7)' Here J is the metric Jacobian, which can be
interpreted as the volume, and x 7J ' etc. are parts of the surface normal vector. The
details can be found elsewhere. After the discretization in the transformed plane with
the corresponding (properly discretized) metric formulations, the resulting discretized
equations agree with discretization derived by the "Finite Volume" method.
In the "Finite Element" approach, the domain 0 is subdivided into finite elements and the solution vector Q is interpolated using (linear) shape functions. With
a weighted residual statement and by use of the Gauss theorem the conservation
equation can be cast in an approximate form, e.g. as in [20], [21].

~Q=
M k J

~t

JF:-dOJF:n-Ndf
ax'

aNj

(17)

with the finite element matrix M jk For further details see the related literature.
Inspecting the three formulations in more detail, it can be seen that the Finite Element
formulations for conservation laws agree exactly or approximately with those from the
Finite Volume or Finite Difference discretizations, see e.g. [21]. It means the different
ways of conservative discretization result in equivalent formulations. Remarkable
differences for spatial discretization can occur by different arrangements of the control
volume, by the different numerical flux formulations, and by the different updating of
the cell-interface fluxes.

42 Computational Methods in Hypersonic Aerodynamics


Numerical flux formulation
The discretization of the conservation equations for a small control volume results in
a system of difference equations for the rate of change of the variables balanced by
the normal fluxes over the cell interfaces. The variables used in the time derivatives
usually are volume-averaged values, whereas the fluxes in the steady-state operator
need cell interface values.
For the sake of simplicity the 1-D conservation equations will be used in the
following discussion. In a common conservative difference form, they read for a grid
point "i":
Si+I/2 - Si-l/2
~x

(18)

Herein ~Q / ~t is the discretized time derivative, defined later, and Fi 1/2 and Si1/2
are the numerical flux functions for the corresponding inviscid, and the viscous fluxes
at the cell interfaces.
The viscous fluxes are updated by means of central differences O(~X2), which
correspond to the elliptic nature of the viscous effects. Then a viscous term of the
form S = I1Ux may be written as

(19)
Cross derivatives can be treated in a similar way.
The numerical formulation of the inviscid fluxes has a great influence on the
properties of the solution method, since they contain the essential information of wave
spreading. These 'terms are strongly nonlinear and are mathematically more difficult
than the viscous terms. Therefore most schemes for the Navier-Stokes equations differ
by the formulation of the inviscid flux terms (Euler solver).
The numerical formulation can be divided into two parts: the evolution of the
numerical flux function at cell faces, which should be a consistent approximation of
the physical fluxes, and the projection of the volume-averaged variables to the cell
faces for updating the fluxes.
In the "projection-evolution" approach, sometimes called "MUSCL" approach,
(van Leer [22]), the basic variables, e.g. Q are extrapolated from both sides to the
cell interface. With these values a common numerical flux formulation is formulated.
In the "evolution-projection" approach, the "non-MUSCL" approach, the flux
functions are formed on the grid points, where the variables are stored, thereafter the
fluxes are projected to the cell interface. This approach is used e.g. for the "cellvertex" schemes or in the modified flux approach [23]. Also mixed forms of both may
be possible. The numerical flux functions become different for the two approaches, in
particular for higher order accurate schemes. However, which of the approaches is to
be prefered cannot be decided uniquely, discussions about that are made e.g. by Yee

[24].
Evolution of numerical fluxes
The evolution of numerical fluxes can be considered as the solution of a Riemann
problem at the cell interface. Within each cell there exists an averaged value of the

Computational Methods in Hypersonic Aerodynamics

43

conservative variables. The values at the interface to the neighbouring cell result in
a jump when they are extrapolated from the left and the right cell-averaged values.
According to the theory of the nonlinear hyperbolic Euler equations this jump of
the values generates a local Riemann problem, whereby information is transported
forward and backward by the different gasdynamical waves and shocks. The solution
of the local, exact Riemann problem results in Euler solvers of the Godunow-type [25],
[26], which describe very accurately the wave phenomena, but these methods are very
expensive. If the jump at the interface is considered to be weak, the Riemann problem
can be solved approximately using the characteristic solution of the Euler equations.
A large number of numerical flux formulations, called approximate Riemann solvers
can be derived with this assumption. Widely used approximate Riemann solvers are
the schemes of van Leer [27], Roe [28], Osher, Chakravarthy [29], Steger and Warming
[30], and many others.
The derivation of the approximate Riemann solvers starts from the 1-D conservative form

(20)
The characteristic form of the Euler equations is derived by diagonalization of the
flux Jacobian A =
which results in an uncoupled system of characteristic Euler
equations. These equations describe the information transport along the characteristic
directions.
(21)

M,

The diagonal matrix A


diag(Ai) = T- 1 AT is formed with the eigenvalues Ai =
(u, u a). The relation between the conservative variables Q and the characteristic
variables W is given by the eigenvector matrix T, i.e. T dW = dQ .
For subsonic flow, i.e. -a ~ u ~ a, the eigenvalue matrix A can be split in a
positive and a negative part, i.e. A = A+ + A-. Then the characteristic form results
m:

(22)
An upwind difference formulation of this equation for the cell "i" with the faces i 1/2
IS:

~W

~t

A+

+ ~x (Wi~I/2 -

A-

Wi::. 1 / 2 )

+ ~x (Wi+1/ 2 -

Wi=I/2)

=0

(23)

Herein W are the backward or forward interpolated variables according to the sign
of the eigenvalues. In the simplest (first order accurate) case they are Wi~I/2 = Wi

W;+1/2 = Wi+l .
The desired conservative difference form can be constructed from the characteristic form by assuming a locally constant characteristic field. This is achieved by
multiplying the characteristic system, Eq. (23) with the eigenvector matrix T, which
results in a conservative form:

and

(24)

44 Computational Methods in Hypersonic Aerodynamics


where Fi 1/2 is the numerical flux function, which includes the characteristic informa
tion.
Using different splitting of the characteristic system, different updating of the cell
interface values, and different approximate, conservative back transformation many
of the different numerical flux functions can be derived from this formulation.
In the following some typical approaches for approximate Riemann solvers will be
discussed briefly.
Flux-vector splitting methods can be derived by a direct application of the conservative back transformation to Eq. (23) and defining A = T A T-l, and F = A Q.
Then the split numerical flux results in
(25)

The flux conservation requires that F = F+ + F-. The values Q are the cell
face values extrapolated from left and right. In the literature there are different
formulations for flux-vector splitting.
The flux-vector splitting by Steger, Warming [30] uses the eigenvalue splitting
A = 1/2 (A IAI). Then the numerical fluxes are defined by:

(26)
This split flux results in a discontinuous numerical eigenvalue whenever the corresponding physical eigenvalue vanishes, van Leer et al. [31]. Therefore it causes a
finite dissipation for steady waves, favorably for strong shocks, but too dissipative for
viscous computations. This property is utilized for a robust hypersonic code by Eberle
et al. [32] using a blended formulation of Steger Warming fluxes and the characteristic flux approach. Furthermore the eigenvalues of Steger Warming fluxes change
discontinuously near the sonic point, which reduces the rate of convergence, as found
in test calculations.
The flux-vector splitting by van Leer [27] avoids the latter drawback of the Steger,Warming fluxes by defining the split fluxes as polynomials of the Mach number,
such that they are smooth near the sonic point. Since one eigenvalue vanishes over
a steady shock, its representation becomes very sharp. Furthermore this splitting, if
applied in implicit schemes, results in very efficient, diagonal-dominant solution methods, e.g. Thomas and Walters [33], Schroder and Hanel [34]. However, this splitting
formulation also shows a remaining splitting error for steady tangential discontinuities.
This drawback influences strongly the accuracy in viscous layers, as shown by Hanel,
Schwane [35]. Therefore a number of modifications was developed and reported e.g.
by van Leer [36] and by Hanel, Schwane [35J. Using these modifications the accuracy
in viscous flows could essentially be increased, as shown by Hanel et al. [35], [37].
The flux-difference splitting method Roe [28], can be derived from Eq. (23) by using
the eigenvalue splitting A = HA IAI). Rearranging Eq. (23) for A and IAI and
applying the conservative back transformation, the numerical flux component Fi+1/2

Computational Methods in Hypersonic Aerodynamics 45


of the flux-difference splitting results in:

-.

_!

F'+l/ 2 - 2 (F(Q;+1/2)

- T--1 (Qi+l/2
+
+ F(Q;+l/2))
+ !2T- IAI
-

Qi+l/2)

(27)

The matrices f' and A are evaluated with Roe averaged values Q for the cell interface,
which shows an improved shock representation.
The eigenvalues are correct represented by this scheme, even if one eigenvalue
vanishes in A. Just this violates the entropy condition in form of expansion shocks
and reduces the robustness of schemes. Following Harten [23] the eigenvalues are
bounded by a so called entropy correction 8, which is a small number of order of
0(10- 1 ), scaled by a typical velocity. This correction can be used in different forms,
e.g.:

(28)
Effects of the entropy correction 8 in viscous flow solutions are studied e.g. in [37].
Central fluxes The central schemes can be considered as a special class of approximate Riemann solvers. Then the characteristic range of influence is equally weighted,
independently of the Mach number by assuming A+ = A- = A/2. The numerical flux
at the cell interface results in:

(29)
The central formulation includes the correct eigenvalues of the flux Jacobian A, and
the spatial discretization error is of second order. However since this numerical approach ignores the characteristic range of influence, the numerical solution shows a
poor shock capturing capability. Furthermore the central formulation does not include
dissipative parts (even derivatives) thus any high-frequency error components of the
solution cannot be damped. Therefore artificial damping terms for high-frequency
damping and shock capturing are added [38].
In general, the damping formulations consist of a linear fourth order term d(4) and
of a nonlinear term d(2). Then the numerical flux reads:

Fi+l/2 =

21 ( Fi+1 + Fi ) + di(4)+l/ 2 -

(2)

di +l/ 2

(30)

The high frequency damping term d(4), necessary to smooth errors of short wave
lengths, (e.g. round-off errors) has the form:

d(4) =

(.::(4)

.6.x) .6.x4 . Q
.6.t
xxx

(31)

where 6(4) is a user specified constant.


The shock capturing term d(2) has to suppress oscillations from the nonlinear
terms in particular for shocks.

(32)

46 Computational Methods in Hypersonic Aerodynamics


The damping is controlled by the nonlinear coefficient c(2), which is proportional to a
constant, multiplied by the normalized second derivative of the pressure. In smooth
regions of pressure the term becomes very small, but it is of O(~x) in regions of
strongly varying pressure.
Refinements were given by various authors, e.g. by Eliason, Rizzi [39] for applications to hypersonic flows, formulations and further references for damping terms in
viscous flows can be found e.g. in [40], [41].
Multi-dimensional Riemann solvers should be mentioned here as an important new
tendency for Euler solvers. Whereas the approximate Riemann solvers, as discussed
above, are derived from a one-dimensional Riemann problem, the multi-dimensional
Riemann solvers take into account multi-dimensional wave propagation and remove
the drawback of mesh dependence of the usually one-dimensional Riemann solver.
Examples and analysis of multi-dimensional Riemann solvers with further references
are given e.g. by Deconinck [42], by Powell, van Leer [43], Rumsey, van Leer, Roe
[44].
Extension of the flux functions to higher order accuracy
The flux-vector and flux-difference splitting schemes are first order accurate in their
original form with the left and right values Q used as Qt+1/2 = Qi and Qi.tl/2 =
Qi+l. But higher order accuracy is required for upwind schemes when applied to
practical computations. Two widely used ways of extensions to higher order accuracy
shall be mentioned here, which are the MUSCL extrapolation (by van Leer [22]), and
the modified flux approach by Harten [23].
MUSCL approach In the MUSCL approach by van Leer [22], the variables Q are
extrapolated higher order accurate to the cell interface and then substituted in the
numerical flux formulation. The MUSCL extrapolation is based on a general polynomial for the forward Qt+l/2' and backward extrapolated values Q+1/2' which can be
written as:
(33)
Qf+l/2 = Qi + 1/4 <Pi ((1 + I\:)~ +Q + (1 - I\:)~ -Q)i

Qi+1/2 = Qi+l - 1/4 <Pi+l ((1

+ I\:)~ -Q + (1 -

I\:)~ +Q)i+l

with ~ +Qi = Qi+1 - Qi and ~ -Qi = Qi - Qi-l.


Herein <P is a limiter function, and I\: is the discretization factor. For <P = 0 it
yields a first order upwind scheme O(~x), with <P = 1 the scheme becomes higher
order accurate, at least second order. In more detail the scheme is then central O(~X2)
with I\: = 1, upwind-biased O(~x3) with I\: = 1/3, upwind-biased O(~x2) with I\: = 0,
and fully upwind O( ~X2) with I\: = -l.
The limiter functions <P are elements of the numerical flux, which limit the higher
order extrapolation to suppress numerical oscillations. In the present definition they
vary between 0 and 1, controlled by the solution itself. Such limiter functions can be
constructed from the theory of almost monotonic solutions (TVD-theory) and applied
approximately to hyperbolic systems in multi-dimensions.

Computational Methods in Hypersonic Aerodynamics 47


A large number of limiter formulations can be found in the literature, e.g. the
Albada limiter [45], the van Leer limiter [22], or the minmod limiter [46]. Comparisons
of different limiters are given e.g. by Sweby [47] or by Vee [24].
Common to all of these limiter formulations is their reaction on the changes of
the local gradient of the variables, expressed by a function r.pi = r.pi(~ +Qi , ~ -Qi, ... ).
The general behaviour ofthe limiters is given by 1-r.p rv IQxx/ Qxl, which can be shown
by expanding the differences in the limiter function. Thus in regions of weak changes
of Q the limiter remains nearly one, but for strong changes the value decreases, the
scheme reduces the accuracy, and stronger numerical dissipation is generated.
The MUSCL extrapolation can be applied to flux-vector splitting as well as to
flux-difference splitting, if they are expressed by the left/right values Q.
Modified-flux approach The modified-flux approach by Harten [23], improved by Vee
[24], is used to increase the accuracy of the flux-difference splitting schemes. The first
order flux-difference splitting, Eq. (27) with Qt1/2 = Qi and Qi+1/2 = QiH consists
of a central flux term of O(~X2) and a first order upwind term. To compensate the
first order term an additional term is added to Eq. (27), such that the scheme becomes
second order accurate in smooth regions. Near extremes or shocks the additional flux
vanishes, which retains the first order upwind term for capturing the shock. The flux
difference splitting for the modified flux approach is usually written in the form:
-

Fi+1/ 2 = 2" (FiH

1-

+ Fi ) + 2"T <I>i+1/2

(34)

The second term corresponds to the modified Roes upwind term in Eq. (27), T is the
eigenvector matrix and <I> contains cell face differences of the characteristic variables
controlled by flux limiters. Examples for this flux splitting are e.g. the "symmetric"
and "upwind" TVD schemes of Vee and Harten. A detailed report about the different
formulations is given by Vee in [24].
The preceding consideration of different Euler solvers covers only a small part of
existing methods, which are in their details so manifold as the authors are. Therefore
critical studies and comparison of the different flux formulations are necessary, in
particular for application to viscous flows. Much information can be found in Vee's
paper [24]. Analysis of different flux algorithms were carried out e.g. by Grossman
and Walters [48], by Montagne, Vee and Vinokur [49] and by Kroll et al. [50]. The
behaviour in viscous solutions were studied e.g. by van Leer et al. [31] and Hanel et
al. in [51], [35], [52], [37].
METHODS OF SOLUTION
Commonly the time-dependent Navier-Stokes equations are used for solutions of unsteady, and stationary flow problems as well. The advantage of using the timedependent equations for steady-state computations is that the initial-boundary value
problem remains parabolic or hyperbolic in the time-space plane, independent from
the Mach number range. Therefore one and the same numerical method can cover

48 Computational Methods in Hypersonic Aerodynamics


a broad range of applications. Thus the methods of solution for the time-dependent
equations correspond in principle to the methods of integration in time. These methods can be classified by their different properties.
The time integration can either be based on explicit methods or on implicit methods.
The explicit methods have the simpler algorithms, only the steady-state operator
has to be evaluated from the known initial state, the new state can be computed
decoupled for each grid point. The explicit methods are well suited for structured and
unstructured grids as well. Vectorization and parallelization of the algorithm is much
simpler. However the numerical time step of an explicit method is restricted by the
numerical stability, depending on the space step divided by the fastest gasdynamical
wave speed, and in viscous flows on the cell Reynolds number. For this reason the
explicit schemes become inefficient for stiff equations, when very different scalings
have to be resolved.
The implicit schemes have the advantage of being unrestricted stable, or at least
allow a much larger time step than an explicit scheme. But the computational work
per time step is much higher due to the inversion of large solution matrices. The
recursive structure of the inversion algorithms is a handicap for vectorization and
parallelization of implicit methods. But nowadays implicit methods can be vectorized
in the same high degree as for explicit schemes with appropriate algorithms, improved
compilers and sufficient computer storage.
Time integration methods can also be divided in classes of schemes, which are
combining space and time integration, and such where space and time integration is
independent.
In schemes combining space and time integration an additional term is added
to the spatial discretization for increasing the temporal accuracy and stabilizing the
scheme. Due to this term these schemes yield steady-state solutions that depend on
the time step. To this class of schemes belong the Lax- Wendroff type schemes. Most
popular is the predictor-corrector scheme of McCormack [53], which was applied e.g.
to complex hypersonic flow by Shang and Scher [54].
Schemes of independent space and time integration are the most used types of
schemes at present. The major advantage is that the steady-state solutions become
independent of the solution method and thus efficient convergence-accelerating means
can be used for steady-state calculations. Typical members of this class are the
explicit Runge-Kutta type methods and the many implicit schemes based on corrector
formulations.
Another classification of solution methods can be made for time accurate schemes
and such methods for steady-state solutions only.
Time-dependent flow problems require a discretization which is consistent and
stable in time and space. Additional care has to be taken for the temporal accuracy,
which is influenced by the discretization errors in time and space as well. For timeaccurate computations the explicit scheme results in a tolerable time advance, as long
as only one time scale is present in the physical problem, e.g. the gasdynamical scaling
in inviscid solutions, just defining the stability restriction. The time step becomes very

Computational Methods in Hypersonic Aerodynamics 49

small and the time accurate computation becomes expensive if additional smaller scale
lengths have to be resolved, like in viscous flows at high Reynolds numbers or in stiff
chemical rate equations. For this reason time accurate computations with an explicit
scheme may be still efficient for inviscid flows and viscous flows at moderate Reynolds
numbers, but for high Reynolds numbers some other ways, like implicit schemes or
multigrid methods should be taken into consideration.
For steady-state computations the transient solutions have no meaning, as long as
the steady-state solution will not be influenced by them. Therefore consistency and
accuracy in time is not required, the method of solution (in time) can be chosen e.g.
for an optimal rate of convergence. This results in classes of pseudo time-dependent
methods which correspond to iteration schemes. Herein different acceleration strategies for improving the convergence can be applied, like e.g. Gauss-Seidel relaxation,
Lower-Upper matrix decomposition, local time stepping or the multigrid methods.
The methods of solutions and their variants are numerous, therefore in the following
section only some of typical methods of solution for the Navier-Stokes equations will
be discussed. For the discussion the two-dimensional Navier-Stokes equations are
considered. The treatment of the 3-D equations is analogous. The equations in
conservative form, written for a curvilinear coordinate system, let's say (~, T/, t), read:
(35)
Details of the transformed equations can be found elsewhere. With a conservative
discretization in a way as described above the approximated Navier-Stokes equations
(35) read:

!:lQ
!:It

+ Res(Q) =

(36)

where !:lQ /!:It is the discrete time derivative, and the residual Res (Q) corresponds
to the discretized steady-state operator:

(37)
The fluxes F, S, etc., are the numerical fluxes, upwind or centrally discretized, and
the difference operators be , bTl mean:

Linear multistep methods for the discretized conservation equations were investigated
by Beam and Warming [55]. Practical applications are restricted usually to implicit
one- and two-step methods. A general method of solution, based on two time levels,
may be written as:

Qn+!

Qn

~--~
!:It

+ (1- 8) Res(Qt + 8 Res(Q)n+l

= 0

(38)

50 Computational Methods in Hypersonic Aerodynamics

For = 0 the scheme is explicit of O(~t), with = 1/2 the implicit Crank-Nicholson
scheme, O(~t2), and with e = 1 the implicit Euler-backward scheme is achieved.
Implicit schemes
The formulation of an implicit scheme requires time linearization of the unknown
fluxes in Res(Q)n+l. This yields:

Res(Qt+ 1 =

Res(Qt+~t.aa Res(Qt+ =
t

Res(Qt+ a ResYJ) (Qn+l_Qn) (39)


aQ

The time linearization of the components of Res( Q)n+l result in the implicit scheme:
I

[~t

+ e {8eA- + 81)B- -

with ~Qn = Qn+l - Qn, and

Re (8 eC

A, B, 6, b

~
+ 81)D)}]n
~Qn =

-Res(Qn)

(40)

are the Jacobian of the corresponding fluxes

F,e,s and T.
The discrete steady-state operator on the right hand side (RHS) is usually discretized higher order accurate with a numerical flux formulation, as described before.
Thus the RHS determines the spatial accuracy of the solution.
For time-accurate solutions, the implicit operator on the left hand side (LHS) has
to be developed consistently to the RHS to achieve time accuracy. Then the LHS
results in a penta-diagonal block matrix system for each direction.
For steady-state calculations, i.e. RH S ---t 0, consistency in time is not required,
and therefore the LHS can be manipulated to achieve faster convergence. To reduce the computational work, and to increase the diagonal-dominance of the solution
matrix three point stencils are used for the LHS. Thus only systems of tridiagonal
block matrices have to be inverted. Such approximations can be achieved in upwind
methods by using first order upwinding for the LHS, partially with additional simplifications (e.g. spectral radius instead of complete eigenvalues [56]). This idea is also
applicable to central schemes as shown e.g. in [56] or special damping terms are used
there.
Even with such approximations for the LHS, the implicit scheme, Eq. (40) requires
the inversion of a large system of difference equations. In general the direct inversion
of the solution matrix for more than one dimension is too expensive, therefore in
most cases the matrix is solved in simpler, approximate steps, using factorization or
relaxation methods.
Approximate Factorization method (AF) The Approximate Factorization method
(AF) by Beam and Warming [55], was the most used implicit scheme in the seventies. The solution matrix is split in one-dimensional operators, which are sequentially
inverted by a Gauss elimination method. For the present example, Eq. (40) with
= 0, the solution steps of the AF method are:

[1+~t8e(A- ~e6)] ~Q
[I

+ ~t 81)(B - ~e

b)] ~Qn

(41)

Computational Methods in Hypersonic Aerodynamics 51


This method enables the inversion per time step in a non-iterative way, which is an
interesting aspect for time-accurate computations.
The efficiency for steady-state solutions is restricted by the fact that the time
step is limited to Courant numbers of 0(10) due to the factorization error in 2-D, and
in 3-D the method becomes even unstable. However with appropriate formulations
of the implicit operator successful algorithms were achieved, e.g. the ARC codes by
Pulliam and Steger [57], the diagonalized AF method by Chausee and Pulliam [58],
and in the ADI method by Briley and McDonald [59]. The factorization method is
used e.g. Lombard et al. [60] for upwind schemes.
Lower-Upper factorization methods (LU) In these methods the unfactored implicit
operator is approximated by two factors Land U, representing a lower and an upper
triangular matrix, and by a diagonal matrix D. For the present example the scheme
would read:
(42)
Herein L contains all the unknowns at points (i - 1,j),(i,j),(i,j - 1), and U the
values for (i + 1,j), (i,j), (i,j + 1), whereas D collects the values in the point (i,j).
The triangular matrices can easily be inverted and the system solved by forward and
backward sweeping over the integration domain. The LU method requires strong
diagonal-dominance, which can be achieved by first order upwinding on the LHS.
The 3-D Navier-Stokes equations and the k-c; turbulence equations were solved
by Yokota [61] with a diagonally inverted LU factored implicit multigrid scheme. A
numerical study of chemically reacting flows using a LU symmetric successive overrelaxation scheme was carried out by Shuen and Yoon [62].
An application to hypersonic flow is presented by Rieger and Jameson in [56],
where the RHS is centrally discretized and the LHS is made diagonal-dominant by
simple upwinding and solved by incomplete LU decomposition (ILU). Numerical experiments of the author and co-workers have shown a similar convergence behaviour
of the ILU method, used in [56], and of a 5-step Runge-Kutta scheme with implicit
residual smoothing, but more robustness for the ILU method could be stated at high
Mach numbers.
Implicit relaxation methods Relaxation schemes based on classical relaxation methods for elliptic equations (Gauss-Seidel methods) became an alternate, very efficient
way for implicit schemes, since strong diagonal-dominant solution matrices could be
achieved with the development of upwind schemes. The relaxation schemes make
use of an iterative procedure to solve either iteratively the steady-state equations, or
for the time-dependent equations an iterative procedure is used for each time step.
Time-accurate solutions can be achieved, if the iterative solution converges for each
time level. For steady-state solutions the time step is often used to increase the
diagonal-dominance in the initial phase, later the time step is increased (to infinity)
with decreasing RHS (residual), see e.g. Schroder, Hanel [34].
For the present example, Eq. (40), the iterative procedure from time level n to

52 Computational Methods in Hypersonic Aerodynamics


n+1 reads
( 43)
where the superscript v is the iteration index and t:~iJ" = Qn+I,"_Qn. The iteration of
equation (43) is performed by either a collective point or line Gauss-Seidel relaxation
in alternating directions. The iterative procedure is stopped if maxl~Q"+I_~Q"1 ~ c
where c is a small number. If the solution matrix is sufficiently well-conditioned and
diagonal-dominant the resulting solver becomes very robust and efficient for steadystate solutions since the time step is indeed unrestricted.
The use of relaxation methods has proved valuable in hypersonic flows, where
robust behaviour of the methods is required. Results of hypersonic calculations with
relaxation-type methods are presented e.g. by Candler and MacCormack [63], Lacor
and Hirsch [64], Schwane and Hanel [35], Riedelbauch and Brenner [65], Schroder and
Hartmann [66].
Explicit Methods
Among the numerous explicit methods for the solution of conservation equations, the
Runge-Kutta time-stepping schemes are the most popular member of explicit schemes
at present.
Runge-Kutta time-stepping schemes Considering the equation (36) as a semi-discrete
approximation of the time-dependent Navier-Stokes equations, the time discretization
can be carried out as a sequence of intermediate steps in the sense of the classical
Runge-Kutta method. At present, a version of the Runge-Kutta method, as published
by Jameson [67], is widely used for solutions of the Euler and Navier-Stokes equations.
This version requires minimum computer storage, but can be extended only to second
order accuracy in time. In the original papers central differencing with artificial
damping was used, but also upwind schemes are applicable.
For aN-step Runge-Kutta method the scheme for equation (36) reads

-0;/

Qn+I

Q(O)

~t

Res( Q(I-I))

+ ~Q(I)

( 44)

= QN

In calculations 3, 4, or 5-step Runge-Kutta schemes are employed, with a theoretical


maximum Courant number of N - 1. The upwind scheme results in a slightly lower
Courant number but with a better high frequency damping, which is of advantage for
the multigrid treatment.
To accelerate the convergence to the steady-state solution local time steps can be
used which are dictated by the local stability limit and constant Courant number. The
local time stepping allows a faster signal propagation, and thus faster convergence.

Computational Methods in Hypersonic Aerodynamics 53

A second acceleration technique is the implicit residual smoothing [67], for which
with appropriate smoothing coefficients the CFL number can be increased by a factor
of two to three.
A further, now widely used acceleration technique, is the multigrid method, as
proposed by Jameson [67] for the Runge-Kutta scheme. More details about multigrid
are given in a later section.
Applications of the Runge-Kutta method to the 3-D Navier-Stokes equations are
reported e.g. by Vatsa and Wedan [40], Cima and Yokota [68], RadesPiel' Rossow
and Swanson [41] on structured grids, and by Mavriplis [21] on unstructured grids.
Time-accurate calculations, using the Runge-Kutta method are given e.g. by Meinke
and Hanel [69].
Methods on unstructured grids
Finite Element as well Finite Volume methods on unstructured triangulated grids
enable geometrical flexibility and adaptive meshing for a high degree. Such methods
have proved valuable for solutions of the Euler equations in complex 2- D and 3-D flows,
e.g. [70], [71], [72]. The same concept applied to the compressible N avier-Stokes
equations for high Reynolds numbers has been not so far developed as for inviscid
flows. The major reason is the presence of very different scalings in viscous flows.
This has the consequences of higher grid resolution, the directionality of gradients
in viscous layers and the corresponding deformation of grid cells. Furthermore the
implementation of frequently used algebraic turbulence models is more difficult than
in structured grids. About the latter item an approach can be found in [73]. For
these reasons attempts are made to combine structured meshes in viscous zones with
unstructured meshes outside. Computations on such hybrid meshes are reported e.g.
in [74], [75]. Another approach is to use structured meshes in one direction normal
to the wall, where viscous terms are important, and in the other directions of a 3-D
grid unstructured organization [76] is used.
Despite of the difficulties mentioned above (and because of the advantages against
structured meshes) the number of computations for the Navier-Stokes equations on
fully unstructured meshes shows an increasing tendency. Navier-Stokes computations
on unstructured grids are shown e.g. in [21], [77], [78]. A large number of 2-D and 3-D
hypersonic applications of finite-element methods for reacting and non-reacting flows
were presented by Periaux [79] and Mallet [80], summarizing the works of INRIA and
AMD- BA in France.
For the solution of the Navier-Stokes equations on unstructured grids nearly all
of the numerical flux approaches were used as known in structured grids. Typical
approaches are central Runge-Kutta scheme with multigrid, e.g. Mavriplis [21], central
two-step Lax-Wendroff scheme, e.g. [78], Roe's flux-difference splitting, e.g. [81], van
Leer's flux-vector splitting, e.g. [82], TVD Osher scheme, e.g. [83], flux-corrected
transport approach, e.g. [78]. The methods of solution are explicit as well as implicit
schemes, details can be found in the related papers.
Summarizing one can say that methods on unstructured grids for the N avierStokes equations have a great potential for geometrical complex flow problems and

54 Computational Methods in Hypersonic Aerodynamics


therefore, the number of applications will increase in the near future.
Multigrid methods
The multigrid method is known to be the most efficient method for solving elliptic
partial differential equations. The basic concepts of multigrid methods were formulated by Brandt [84], [85]. Encouraged by this success, attempts were made to take up
the multigrid concept in solution methods for time-dependent, parabolic or hyperbolic
problems. Here, of very great interest are the numerical solutions of the conservation
equations, which require a large amount of computational work for practical problems. Thus, to reduce the computational work, a number of investigations were made
to incorporate the multigrid concept into existing methods of solution for the Eulerand Navier-Stokes equations as well. Examples for multigrid applications in explicit
and implicit Euler solutions are given by Chima, Johnson [86], Ni [87], Jameson [67],
and by Hemker [88], Mulder [89] and others. Navier-Stokes applications can be found
e.g. in the paper of Shaw, Wesseling [90], Thomas et al. [91], Schroder, Hanel [34],
Hemker and Koren [92], Radespiel, Swanson [93], Jameson, Siclari [94], Vatsa and
Wedan [40], Abid and Vatsa [95], Radespiel, Rossow and Swanson [41], Mavriplis
[21], Yokota [61], Gerolymos [96], Meinke and Hanel [69], [97]. An overview of the
present state of multigrid is given on special multigrid conferences, e.g. the "Copper
Mountain Conference on Multigrid Methods", Copper Mountain, Colorado, and the
"European Conference on Multigrid Methods", Bonn, Germany.
The basic multigrid concept for nearly all of the applications is the Full Approximation Storage concept (FAS), as proposed by Brandt [85]. This concept is developed
for the solution of nonlinear equations, and therefore well suited for the conservation
equations. A requirement for the application of multigrid formulations is the property
of smoothing the high frequency error components by the scheme. This requirement
is satisfied in principle by the use of an upwind scheme. The FAS multigrid concept
can be applied to implicit and explicit methods of solution. Applications of multigrid
methods to an implicit relaxation scheme and to an explicit time-stepping scheme
were studied e.g. by the author in [34] and [97]. For the implicit relaxation method
the multigrid method is used to accelerate the iterative matrix inversion each time
step [34]. Thus it is part of the relaxation procedure for each time step. For the
explicit scheme the multigrid concept is applied in space and time as well. It directly
influences the solution in time, and therefore it can be used for the solution of steady
and unsteady flow problems [97].
The gain of computation time for multigrid applications to the compressible
Navier-Stokes equations was a factor of two to ten, not so much as the gain for
elliptic equations, and partially for the Euler equations. One reason for that is the
high aspect ratios of computational cells, if computing flows at high Reynolds number. Another handicap for multigrid methods can be stated in strongly supersonic
(hypersonic) flow, where the solution is hyperbolic in space, and therefore it is in
contrast to the "elliptic" multigrid concept. This seems to be the reason why only
few publications are devoted to multigrid methods for hypersonic flows. One example
was presented by Siclari, Jameson [94]. A multigrid algorithm for an explicit Runge-

Computational Methods in Hypersonic Aerodynamics 55


Kutta scheme with local time stepping and residual smoothing has been applied to
the computation of both inviscid, and viscous supersonic/hypersonic conical flows.
With the exception of one case, the multigrid algorithm reduced the computational
time by at least a factor of two, whereby the multigrid gain was somewhat better for
viscous flow (since the explicit non-multigrid scheme slows down due the viscous grid
stretching) .
Nevertheless, the multigrid methods can be a useful tool in hypersonic flow, perhaps not for the supersonic main flow direction, but it can reduce the computational
effort in the cross flow plane where the flow is subsonic.
EFFECTS OF NUMERICAL DISSIPATION
Methods of solution for the Navier-Stokes equations usually are constructed by adding
the viscous terms to a reliable and stable working solution method for the Euler equation. This combination generally results in efficient Navier-Stokes solvers, stable even
for high Reynolds numbers, and well suited for resolving the inviscid flow portions.
However the Euler solver can generate a large amount of undesired numerical dissipation in the viscous layers with strongly changing flow quantities. This dissipation
is superposed on the physical dissipation in viscous layers, and impairs the solution
there.
The aim of the present section is to discuss such of effects of numerical dissipation
in solutions of the Navier-Stokes equations. The basic discretization schemes for the
Euler equations, and inherent sources of numerical dissipation in central and highresolution schemes for the Euler equations will be discussed briefly. The consequences
for solutions of the Navier-Stokes equations are considered then for steady-state solutions; more information, in particular for unsteady flows, is given e.g. in [52].
Numerical damping in solutions of the Euler equations
The quality of a Navier-Stokes solution is essentially influenced by the numerical
damping caused by the discretization of the inviscid terms (Euler equations).
The Euler equations are a nonlinear, hyperbolic system of partial differential equations. Two classes of solution exist, the strong continuous solutions describing the
non-dissipative wave transport along characteristics, and the weak, discontinuous solutions, e.g. shocks. An efficient numerical method of solution should resolve both
classes of solutions sufficiently well, but then numerical damping has to be used.
Discretizing the Euler equations, the resulting truncation error consists of even
and odd higher order derivatives, which cause dissipative and dispersive errors in the
numerical solution. Particularly the dissipative parts (even derivatives) act viscositylike and cause smearing of the solution and artificial vorticity production as well.
Therefore the aim of the numerical approximation should be to minimize the amount
of numerical dissipation.
On the other side a certain amount of numerical damping is necessary to avoid
the accumulation of randomly distributed small errors (e.g. round-off errors) during
the time progress. In general, these errors are distributed with smallest resolvable

56 Computational Methods in Hypersonic Aerodynamics


wave length which is in the order of the step sizes. Therefore the damping terms have
to filter these error components.
A widely used approach for such high frequency damping terms is to use fourth
order differences, which have a sufficient filtering effect and are small of 0 (.6.x 3) in
smooth solutions. These high frequency damping terms are efficient filters in smooth
regions, but fail in regions of captured, stronger discontinuities. Here stronger oscillations are generated by the frequency amplification of the nonlinear convection
terms in combination with the strong changes over a few grid points. To suppress
these oscillations additional shock capturing terms have to be implemented in Euler
solvers. Usually, such terms are based on second order differences with a nonlinear
viscosity-like coefficient, which activates a strong dissipation within the "shock layer",
but vanishes outside in smooth regions.
Both types of damping terms are essential parts of each numerical flux formulation
for the Euler equations. These terms are either internally generated by the scheme,
e.g. in high resolution upwind schemes or have to be added artificially as in central
schemes.
To understand the behaviour of these damping terms, the most important numerical formulations of Euler fluxes shall be considered in the following. For the discussion
only the 1-D conservative equations are used in a discretized form:

(45)
Herein .6.Q;/.6.t is the discretized time derivative, defined by the method of solution,
and Fi 1/2 are the numerical fluxes at the cell interfaces i 1/2.
Central schemes The central schemes are widely used for approximating the Euler
equations. However central differences do not result in dissipative truncation errors,
therefore such artificial damping terms have to be added. As mentioned above, the
damping formulations consist of a linear fourth order term d(4) and of a nonlinear
term d(2). Then the numerical flux reads:
( 46)
The high frequency damping term d(4), necessary to smooth errors of short wave
lengths (e.g. round-off errors) has the form as given in Eq. (31). The shock capturing
term d(2), Eq. (32), has to suppress oscillations from the nonlinear terms in particular
for shocks.
The effect of the damping terms becomes evident by considering the truncation
error. Applying the numerical formulation, Eq. (46), to the linear, hyperbolic model
equation Qt + Qx = 0, the truncation error in space for a central discretization with
damping terms is:

_ ((2).6.X)
.
_ ~ 2.
Ts- c .6.t .6.x Qxx 6.6.x Qxxx

(c

(4).6.x
3.
.6.t).6.x Qxxxx+ ...

(47)

Computational Methods in Hypersonic Aerodynamics 57


Assuming the Courant number ~~ being 0(1), then the artificial damping terms are
of order O(.6.x), and O(.6.x 3 ), respectively.
Upwind schemes These schemes take the advantage of the hyperbolic properties in
form of an approximated discrete Riemann problem, e.g. flux-vector splitting or
flux-difference splitting. In general they are used with higher order accurate upwind discretization. Nearly oscillation-free solutions and sharp shock representation
is achieved making use of the total variation diminishing (TVD) principle. The development of these schemes is connected with Osher [29], Harten [23], van Leer [22]
and many other authors.
The high resolution schemes do not use any added artificial damping terms,
however the corresponding mechanisms for high frequency smoothing and the shock
capturing are implicitly included in the scheme. Two widely used concepts of such
schemes are considered briefly in the following.
In the flux-vector splitting concept, as proposed by Steger, Warming [30] or by
van Leer [27], the flux is split in two parts with corresponding positive and negative
eigenvalues of their Jacobian.

(48)
The variables Q are variables extrapolated from the left (+) and the right (-) to
the cell interface. For the higher order extension of upwind schemes, van Leer has
suggested the so called MUSCL approach [22], where the variables Q are extrapolated
with a higher order polynomial from both sides to the cell interface, Eq. (33).
The damping properties of such an upwind scheme are strongly influenced by
the extrapolation, Eq. (33), controlled by the discretization parameter K, and by
the limiter r.p. This can be demonstrated by means of the spatial truncation error,
analogous to Eq. (47) by applying the extrapolation (33) to the linear model equation
Qt + Qx = 0. The truncation error for Qx yields:

.6. x

Ts

.6. x 2

.6.x 3

= (1-r.p)TQxx-[1-2r.p(1-K)]-6-Qxxx-[3(r.p-K)-(1-r.p)]UQxxxx+ ... (49)

The flux-difference splitting concept, as proposed by Roe [28], Harten [23], Vee [24],
and others, results in a first order upwind scheme, which consists of a central flux
term and an upwind term taking into account the different directions of information.
To extend the flux-difference splitting to second order accuracy the MUSCL principle,
Eq. (33), can be used, or an additional term 9i+1/2 can be added, according to the
modified flux approach, Eq. (34), by Harten [98].

Fi+1/ 2 =

1-

2 (Fi+l + Fi ) + 2T ~i+1/2

(50)

The different formulations for the modified flux are summarized and discussed e.g.
in [24]. With a simplified form of this flux the spatial truncation error Ts of flux
difference splitting, applied to the linear model equation Qt + Qx = 0, can be written
as:

58 Computational Methods in Hypersonic Aerodynamics


The numerical flux formulations for the three schemes, Eq. (46), (48) and (27), are
very different, however their reduction to truncation errors for a linear model equation,
Eq. (47), (49) and (51), show common properties for the different flux approaches.
Each truncation error consists of three parts, a shock capturing term rv ~x Qxx, a
dispersive term rv ~X2 Qxxx, and a high frequency filter term rv ~X3 Qxxxx.
The shock capturing term is controlled by a nonlinear function (pressure dependent term in central schemes or limiters in high resolution schemes). The nonlinear
function reduces the capturing term to higher order accuracy in smooth regions, but
generates first order damping near extremes or shocks. The latter effect results in
a sharp, non-oscillating solution of embedded shocks. In viscous layers with large
changes of the variables this term is also active and impairs the viscous solution.
The high frequency filter term of accuracy O(~X3) is fixed by a user specified
constant in central schemes or by the upwind formulation in high resolution schemes.
It filters the undesired high frequency error components, but in viscous flows with
small viscous scale lengths, it can also suppress small-scale vortical structures.
Numerical damping in solutions of the N avier-Stokes equations
The combination of Euler solver and viscous terms generally results in efficient and
stable Navier-Stokes solvers. But in contrast to Euler solutions, the solutions of the
Navier-Stokes equations contain very different characteristic scale lengths with corresponding different types of solution. Typically are thin viscous layers with continuously, but strongly changing flow quantities, and large nearly inviscid flow portions
with discontinuous solutions. Discontinuities arise from the fact, that the very thin
viscous shock structures cannot be resolved in general. Thus, the solution of the
Navier-Stokes equations requires good shock capturing properties of the Euler solver
and a high accuracy in viscous regions as well. These requirements are often contradictory, in particular when inviscid and viscous terms are nearly balanced. Most of
the Euler solvers work well for inviscid flow problems, mainly due to the implemented
damping mechanisms. However these mechanisms usually cannot distinguish between
strong gradients in a captured shock and in a boundary layer. Thus in viscous layers,
where strong changes of gradients occur, the numerical dissipation can become large
and is superposed to the physical, viscous effects. This fact can impair essentially
the accuracy of the viscous solution. Therefore it is necessary to know the sources
of damping in solvers, and to minimize their effects in viscous layers. An analysis
of these effects is quite difficult because of the complex nonlinear structure of the
equations and their different numerical approaches. A consideration of different Euler
schemes [52] has shown some possible sources of damping. Such sources can be:
different kind of upwinding (central to fully upwind)
limiter functions
artificial damping terms
numerical flux formulations (complete fluxes, flux splitting)
representation of the eigenvalues in numerical fluxes

Computational Methods in Hypersonic Aerodynamics 59


grid stretching and skewing
numerical boundary conditions
In general all of these effects are interacting and thus cannot be considered isolated.
Therefore numerical experiments and comparisons with other results are useful means
to reveal the effects. In the following a number of results will be shown which demonstrate the influence of different numerical Euler scheme.
Considering the classical central schemes, Eq. (46), with linear fourth order and
nonlinear second order terms the investigations have shown a sufficient spatial accuracy for the boundary layer solution. The reason may be that the linear damping
term can be held small per external parameters, and the nonlinear terms controlled by
the curvature of the pressure, remain small in shear layers where the normal pressure
changes are small. However in boundary layers with strong adverse pressure gradients
the deviations become stronger and require a careful analysis.
More attention has to be paid for the upwind Euler schemes within a NavierStokes solver. Such schemes are very well suited for capturing gasdynamical wave
phenomena by using local approximate Riemann-solvers, higher order upwinding, and
TVD flux limiters. The different damping mechanisms are included in the numerical
flux formulation, and therefore they cannot be controlled directly by the user.
To demonstrate basic effects several simpler test cases were calculated. One test
case is the solution of the nonlinear, scalar Burgers equation Qt (Q2/2)x = (IJQx)x,
using the MUSCL approach, Eq. (33) for the discretization. An exact steady-state
solution of this equation with the boundary conditions Q(O,t) = 0 and Q(oo,t) = Qoo
is given by Q(x) = Qoo tanh (Q2,,;;X). Other test cases are e.g. Navier-Stokes solutions for a boundary layer over a flat plate or free shear layers. In some cases the
effects could be demonstrated also for more complex problems.
The influence of the upwind extrapolation was studied by varying the discretization
parameter '" with c.p = 1 in Eq. (33) in solutions of the Burgers equation. The influence of different upwinding becomes apparent in Fig. 3, where the relative error
between the exact and the numerical solution of the Burgers equation is plotted over
x. The tendency is that for smaller values of '" (i.e. more upwinding) the error tends
to negative values, which means the solution smears out more and more, although the
schemes are higher order accurate. Navier-Stokes solutions e.g for the boundary layer
flow confirm this tendency.
The influence of the flux limiters is studied by applying a typical limiter, e.g. of AIbada [45], to the solution of the Burgers problems. The limiters react upon the
curvature by reducing the higher order terms, which leads in smooth regions to an
undesired numerical dissipation. This tendency is clearly shown in Fig. 4, where the
relative error of the scalar Burgers problem is plotted for different schemes without
and with limiter. All the results with a limiter show a larger negative error, i.e. more
dissipation.
This behaviour is reflected in Navier-Stokes solutions, in particular for laminar

60 Computational Methods in Hypersonic Aerodynamics


flows. In turbulent shear layers damping effects from upwinding and from limiters
were found to be less critical [37], because of the much stronger turbulent viscosity.
Nevertheless the limiter can have significant influence on the results for more complex
turbulent flows, e.g. for transonic flow around an airfoil. Fig. 5 presents the pressure
distribution for transonic flow over a RAE 2822 airfoil [37]. The two numerical results
are achieved with a flux-difference scheme in one and the same grid, but using two
different minmod-limiters with w = 2 and w = 1 (Roe limiter). Differences between
the two solutions are obvious near the leading edge, where the suction peak is somewhat degenerated by the less compressive limiter w = 1. This results in a significant
increase of drag by about 15%. It is therefore advisable to use more compressive
limiters for a better accuracy in viscous computations. But such limiters reduce the
robustness, in particular in hypersonic flows. Therefore in some cases less compressive
limiters are used for the transient state.
The formulation of the numerical Euler fluxes has a large influence on the accuracy
in viscous layers. The Euler fluxes represent a composition of information transported along characteristics (eigenvalues of the flux Jacobians) from different directions (signs). Numerically the direction is expressed by left or right extrapolated
variables. Depending on the decomposition (flux splitting) certain eigenvalues may
be not correctly represented by the numerical flux. Especially for shear layers, degenerating to tangential discontinuities for Re -+ 00, it is important to approximate
correctly the transport of entropy and tangential velocity along the linear eigenvalues,
e.g. A = v. For these eigenvalues the flux-difference splitting has shown to be a better
approximation than the flux-vector splitting by van Leer. This fact is demonstrated
e.g. in Fig. 6, where the skin friction coefficient over a flat plate is plotted using
van Leer's flux-vector splitting with second order MUSCL approach, [27], and the
second order flux-difference splitting by Yee, Harten [24]. The results from the fluxdifference splitting compare well with the Blasius solution, whereas the flux-vector
splitting shows remarkable deviations.
The behaviour of van Leer's flux-vector splitting in viscous regions was studied
in more detail in [35], [99]. It could be shown that the upstream/downstream extrapolation for the split fluxes, Eq. (48), generates a flux defect for the momentum,
tangential to the cell interface. But this defect could be essentially removed by applying one-sided upwinding in direction of the normal velocity (to the cell interface) to
split fluxes of the tangential momentum equation. The one-sided extrapolation of the
tangential velocity reduces the splitting error, and makes flux-vector splitting more
consistent with the correct Riemann problem, in particular for tangential discontinuities. The improved accuracy is demonstrated in Fig. 7 for the skin friction over a flat
plate by comparison with the original formulation by van Leer.
The same modification of van Leer's splitting is used in 3- D solutions of the NavierStokes equations for hypersonic flow over a double-ellipsoidal body [100]. Fig. 8a and
Fig. 8b show vectors of the local skin friction on the body using the same, relatively
coarse mesh, but computed with the original splitting (left), and with the modification
(right) from [35]. The remarkable difference between both figures is the reattachment

Computational Methods in Hypersonic Aerodynamics 61

zone on the "canopy" , which could be resolved by the computations using the modified
fluxes in agreement with experiment.
Another significant influence of the numerical flux formulation in context with van
Leer's splitting was noticed in supersonic blunt-body calculations where significant
deviations of the computed wall temperature were found in regions of large Mach
number gradients [101]. Investigations have shown that this effect is mainly caused
by the non-preservation of the total enthalpy H t using the original split energy flux.
This effect could be removed substantially by using an alternate split energy flux [101]:
(52)
where the enthalpy H t is transported as a whole by the split mass fluxes Ft It has
the advantage of being simpler and of being generally valid also in the case of real
gases. The improvement of the accuracy by this formulation is clearly demonstrated
in Fig. 9 for the wall temperature distribution over a 3-D hemisphere-cylinder body.
The idea of this alternate split energy flux, Eq. (52), was extended to generalized
flux-vector formulations in [32]. Using this change of split energy flux a better preservation of total temperature could be stated.
APPLICATIONS TO HYPERSONIC VISCOUS FLOW
Typical properties of hypersonic flow
Viscous hypersonic flow in the continuum regime are characterized by some typical
features, which are not or not so strongly present in other flow regimes, like subsonic
and transonic flows. Therefore numerical methods have to satisfy some special requirements for simulating such flows. Typical properties of hypersonic flow are:
Strong shocks are appearing. The strongest shock, the bow shock, can either be captured or fitted by numerical methods. Imbedded shocks are weaker and are captured
in general. Therefore good shock capturing methods are required.
Strong expansions can occur, e.g. by transient shock movements e.g. in the initial
phase, or stationary from the stagnation point around the nose. Expansions are continuous solutions, but with a rapid decrease of temperature and density. The decrease
can result in very small values or even in negative values for temperature or density for
accelerated numerical methods during the non-physical transient phase. Very small
or negative values mean loss in convergence or even break-down for the numerical
method.
Hypersonic vehicles are blunt bodies or airplanes of low aspect ratio. Therefore the
flow is strongly threedimensional and thus powerful numerical methods in general coordinates are required.
The 3-D, viscous flow is rather complex, even when the boundary layer theory is valid.
But in many cases this theory cannot be applied since strong viscous-in viscid interactions and strongly separating flow for high angle of attack is occurring. To resolve the
viscous phenomena, fine grids and highly accurate methods of solution are required.

62 Computational Methods in Hypersonic Aerodynamics


At lower altitudes of the flight path of a reentry vehicle large portions of the flow are
turbulent. But turbulent closures and transition criteria are not very well developed.
This uncertainty influences strongly the prediction of friction drag and heat flux.
At higher altitudes the density decreases and with it the Reynolds number. The flow
becomes laminar, which makes the problem of closures much easier. However viscousinviscid interactions and separating flows are more sensible for laminar flow than for
turbulent flow.
With further increasing altitude the Reynolds number becomes small and Mach number high, thus the Knudsen number, I< n rv M a/ Re, increases and rarefaction effects
become apparent. In the near continuum range the Navier-Stokes equations are still
valid. Rarefaction effects may be simulated by changed wall boundary conditions, like
incomplete accommodation of momentum and energy (slip conditions). Gaskinetical
solutions of the Boltzmann equation have to be considered for higher Knudsen numbers up to free molecular flow.
A further aspect of hypersonic flow is the chemical behaviour of air. With increasing
stagnation temperature and decreasing density at higher altitudes additional internal
modes of the molecules and chemical reactions, like dissociation and ionization become excited. If the chemical scaling lengths are much smaller than any geometrical
length, the real gas effects can be considered to be in equilibrium. For this assumption
the conservation equations do not change, only the equations of state have altered.
However, if the chemical scale lengths are comparable with the geometrical one, nonequilibrium effects have to be taken into account. Additional conservation equations
for the species have to be solved, coupled with the usual set of equations. The equatiuns contain chemical source terms, which increase the mathematical stiffness of the
problem.
Numerical methods for hypersonic viscous flow
A consideration of the literature, journals as well as conference reports, shows that
nearly all existing approaches for solving the Navier-Stokes equations are able to deal
with hypersonic flow.
The final quality and convergence of the solution is mainly determined by the grid
resolution and by careful adaption of the algorithm to the requirements of hypersonic
flows. This means work in detail and numerical experience. Therefore only a coarse
classification of the methods of solution can be given.
Methods on structured grids enable the implementation of nearly all discretization
schemes. Structured grids are well suited for efficient implicit methods. But grid
adaption becomes difficult and grid singularities can occur.
Pure explicit schemes for hypersonic viscous flows at high Reynolds number suffer
under the strong stability restriction, and in general they are not so robust for such
flows. Nevertheless hypersonic computations with explicit schemes can be found in
the literature.

Computational Methods in Hypersonic Aerodynamics 63

An example for using an explicit scheme for hypersonic flows is the "historical"
paper by Mac Cormack, 1969, [53], where the well-known predictor-corrector method
was applied to the Navier-Stokes equations. Also nowadays this method is in use
as a reliable and cheap method of solution. A very complex application for that is
given by Shang, Scherr [54], where the hypersonic flow around a complete aircraft was
computed with an algebraic turbulence model included.
Hybrid explicit-implicit schemes are a compromise to avoid the severe stability restriction by an implicit procedure in direction normal to the wall. An example is
given by Riedelbauch, Kordulla [102] for the preceding test problem.
Implicit methods and relaxation schemes are more robust, if strong shocks and very
different characteristic scale lengths have to be resolved, as it is in high Reynolds
number flow or chemical reaction flow. Typical approaches for this type of method
are discussed in a preceding section.
Fully implicit relaxation schemes, based on higher order upwinding and diagonaldominant solution matrices, have shown very efficient convergence properties and
shock capturing capabilities in hypersonic applications. An example of that is the
3-D relaxation method by Schwane, Hanel [35], [99], based on flux-vector splitting.
Results were presented above (e.g. Fig. 8).
A number of hypersonic viscous flow applications are presented in the paper by
Schmatz [103]. The method of solution, called NSFLEX, is based on a third order
accurate local characteristic flux extrapolation, blended with a flux-vector splitting of
Steger-Warming type for capturing of strong shocks. The unfactored implicit scheme
is solved by a point Gauss-Seidel relaxation method. The flux-vector splitting was
reformulated in [32] in a generalized flux formulation, including the modified split
energy flux, as proposed in [101].
The application of flux-difference splitting (Roe-type) to hypersonic flows shows
an increasing tendency. The high accuracy of this splitting and sufficient robustness
can be combined by special implicit relaxation formulations for steady-state solutions. Implicit relaxation schemes for the 3-D N avier-Stokes equations, based on
flux-difference splitting, are described e.g. by Miiller [104] and by Riedelbauch and
Brenner [65]. In [65] comparison was made between the upwind method and a central
scheme including fitting of the bow shock [105]. Nearly the same performance and
accuracy was reported for both schemes, but with a superiority for the upwind scheme
if strong embedded shocks are present. The relaxation scheme in [104] also can be
used in a parabolic space marching manner.
Recently Schroder [66] has presented a 3-D relaxation method for the NavierStokes equations, based on flux-difference splitting, and demonstrated the capability
to deal with very complex hypersonic, laminar and turbulent flows. As an example
from [66], Fig. 10 presents the results for the laminar flow during the stage separation
of a two-stage space vehicle at M a oo = 6. and Re/m = 4.2 x 10 5 A two block
mesh was used with 51 x 139 x 90 and 40 x 139 x 30 grid points. The Fig. 10 shows
the corresponding surface mesh and the computed lines of constant density for the

64 Computational Methods in Hypersonic Aerodynamics


symmetry plane.
An example of an implicit method with an explicit-like solution procedure is presented by Rieger and Jameson [56]. The steady-state Navier-Stokes equations are
discretized centrally with artificial damping terms. The unfactored implicit operator
becomes diagonal-dominant by a simplified splitting and solved approximately by LU
decomposition. In this paper the flow around a HERMES space shuttle was computed
for viscous, and inviscid flow as well. As an interesting example, the Fig. 11 shows
the iso-mach lines in a one cross section of the shuttle for both types of flow. The
comparison reveals the different formation of the vortical flow in particular in the
vicinity of the body.
Space marching methods allow computations in a very economical way for flows which
satisfy the presumptions of attached shear layers and supersonic outer flow in mean
flow direction. The space marching can be carried out with the stationary parabolized
Navier-Stokes (PNS) equations or with the time-dependent thin-layer Navier-Stokes
equations.
Commonly the stationary PNS-equations are used for space marching. Using
these equations special manipulations of the pressure in the subsonic viscous layers
have to be implemented [3], [2]. Applications can be found for the computation of
supersonic, and hypersonic flow over complex geometries or complex physical flows.
An example for the latter may be the paper by Prabhu, Tannehill [106], where a
PNS-code is presented, solving the conservation equations for a multi-component,
chemically reacting gas mixture.
In interesting contribution is given by Power, Barber [6], where a comparison is
made between solutions of the Navier-Stokes equations (NS), the parabolized equations (PNS), and the so called reduced Navier-Stokes equations (RNS). The RNS
equations correspond in their form the PNS equations, but the streamwise pressure
gradient term was modified to account for the upstream propagation of pressure waves
within the subsonic layers. The RNS solution is somewhat more expensive than the
PNS solution, but much cheaper than the NS solution. Results are given for strong
viscous/inviscid interactions on a compression ramp at Ma=14. Fig. 12 shows a comparison of the skin friction distribution using NS, PNS, and RNS procedures. The
results confirm that the PNS equations are not able to describe strong interactions,
even without separation, whereas the results from the RNS equations are closer to
that of the full equations.
Space marching can also be carried out by solving the time-dependent thin-layer
Navier-Stokes equations. The idea is to use a general time-marching algorithm, but
to converge the solution to steady-state in each cross-section across a space-marching
direction. Then the integration in time for one cross-section has the meaning of
an iteration-like procedure for steady-state calculations. An advantage of modifying
time-marching methods for space-marching is that one general method can be used
for subsonic regions, e.g. near the stagnation point, to obtain the initial conditions
for space marching in supersonic flow further downstream.
An example for such a space-marching method is published by Menne and Weiland

Computational Methods in Hypersonic Aerodynamics 65


[4]. In this paper a flux-difference scheme is solved explicitly in each cross-section.
The method was applied to complex hypersonic flow problems, e.g. to the flow over
a two-stage transport vehicle. The Fig. 13a gives an impression of the geometry by a
plot of the isobars on the upper surface for this problem. Fig. 13b and Fig. 13c show
a comparison of the isobars in one cross-section computed for inviscid and viscous
flows.
Methods on unstructured grids, Finite element or Finite volume methods enable flexible grid point enrichment in region of strong changes. The problem of grid singularities
does not occur. Higher order upwind methods are hard to implement, but sufficient
accuracy can be achieved by grid point enrichment. Unstructured grids are not well
suited for implicit methods, therefore explicit methods are preferred. In some cases
hybrid structured-unstructured grids are used, e.g. by Hassan et al. [76] to avoid the
severe stability restriction in zones of small cells by solving implicitly in one structured
direction (mostly normal to the wall).
A large number of 2-D and 3-D hypersonic applications of finite-element methods
for reacting and non-reacting flows were presented by Periaux [79], summarizing the
works of INRIA and AMD-BA in France.
CONCLUSIONS
The preceding chapters have shown a variety of numerical methods and problems for
hypersonic, viscous flows. The consequences can be summarized as follows:
N avier-Stokes solutions for hypersonic flows require a robust and highly resolving
Euler solver to handle the strong gasdynamical wave phenomena. The majority of
methods uses upwind, high resolution Euler schemes, based on the TVD principle.
The strongest shock, the bow shock, can be calculated as a discontinuity by a shock
fitting procedure, this gives the correct shock conditions and saves grid nodes in comparison to the capturing. The embedded shocks are usually weaker and are captured.
Capturing of the bow shock avoids the more complicated fitting procedure. Then
the shock is resolved over several grid points, the number of points depends on the
properties of scheme used, and on the grid arrangement (shock surface orientation).
This "smearing" of the shock can be a disadvantage if smaller scale lengths, e.g. from
chemical reactions, have to be resolved. Similarly important and difficult are strong
expansions, where the thermodynamical quantities, like temperature and density, approach very small values. Numerically, this behaviour can cause severe problems, e.g.
incorrect solutions and a loss of convergence.
The thin viscous layers require a fine resolution and a very accurate algorithm, since
here strong gradients and curvatures are present. Special attention has to be given
to the numerical dissipation of the Euler solver which is superposed the physical dissipation and impairs the viscous flow solution.
The survey has shown that explicit and implicit schemes are used as well. Both types

66 Computational Methods in Hypersonic Aerodynamics


are able to capture the gasdynamical waves. However for flows at high Reynolds
numbers the stability restriction of the explicit schemes reduces significantly the efficiency. This restriction can be partially reduced by some acceleration techniques for
steady-state computations (e.g. local time steps, multigrid etc.) With the increasing
difference of the characteristic scale lengths to be resolved (increasing stiffness), the
implicit methods become more efficient since they are not restricted by the numerical
stability. In particular, in chemical reacting flows where additional small scale lengths
have to be resolved, this fact should be taken into account.
Among the implicit schemes, the number of schemes using an iterative procedure
for the matrix inversion each time step (relaxation schemes), shows an increasing tendency, compared with that using non-iterative methods like approximate factorization
methods. The advantage of the unfactored relaxation schemes is that the time step
is really unrestricted and that different appropriate relaxation techniques can be applied (e.g. point or line relaxations, multigrid). Furthermore most of the relaxation
schemes allow a more flexible organization of the algorithm, either minimizing the
storage requirement to one set of the variables plus geometry per grid point, or if
sufficient storage is available, CPU time can be saved by storing the quantities which
are constant during the iteration loop.
Parallel to solutions on structured grids using finite difference/volume methods a
large number of calculations are carried out on unstructured grids mainly using finite
element methods. The primary advantage of the latter is the geometrical flexibility, avoiding the mesh singularities of the structured grids, and the more economical
mesh arrangement. Adaptive grid refinement by mesh enriching allows an increased
accuracy in critical regions. Drawbacks of the unstructured formulation is the more
expensive calculation and the increased storage requirement per grid node. The unstructured algorithm makes the vectorization more difficult. The solution method
is usually restricted to (decoupled) explicit schemes, thus the schemes may become
inefficient for high Reynolds numbers due to the stability restriction. In future developments the advantages of unstructured and structured grids, and of finite element
and finite difference/volume methods should be combined.
A further important point is the utilization of the computer architecture, i.e. high
rates of vectorization and parallelization are required for the very expensive NavierStokes computations. Hereby the explicit structures are of advantage. However also
most of the implicit schemes are highly vectorizable with more effort (and storage),
in particular for threedimensional algorithms. Parallelization of the algorithms on
multi-processor machines becomes of increasing interest.

Computational Methods in Hypersonic Aerodynamics 67

References
[1] Vincenti, W., Kruger, C.: Introduction to Physical Gasdynamics. Wiley and Sons, Inc., New
York, 1967.
[2] Schiff, L.B., Steger, J .L.: Numerical Simulation of Steady Supersonic Viscous Flow. AIAApaper, No. 79-0130, 1979.
[3] Vigneron, Y.C., Rakich, J .V., Tannehill, J .C.: Calculation of Supersonic Viscous Flow over
Delta Wings with Sharp Supersonic Leading Edges. AIAA-paper, No. 78-1137, 1978.
[4] Menne, S, Weiland C.: Split-Matrix Marching Methods for Three-Dimensional Viscous and
Inviscid Hypersonic Flows. Notes on Numerical Fluid Mechanics, vol 29, Vieweg Verlag, 1990.
[5] Cline D. D., Carey G.F.: Shock Sensitivity in Parabolized Navier-Stokes Solution of High
Angle-of-Attack Supersonic Flow. AIAA Journal, vol. 28, pp. 406-413, 1990.
[6] Power, G.D., Barber, T.J.: Analysis of complex hypersonic flows with strong viscous/inviscid
interaction. AIAA-J., vol. 26, No.7, pp 832-840, 1988.
[7] Srinivasan, S., Tannehill, J .C., Weilmuenster, K.J.: Simplified curve fits for the thermodynamic properties of equilibrium air. NASA RP 1181, 1987.
[8] Liu Y., Vinokur M.: Equilibrium Gas Flow Computations. I Accurate and Efficient Calculation
of Equilibrium Gas Properties. AIAA paper No. 89-1736, 1989.
[9] Bird, R.B., Stewart, W.E., Lightfoot, E.N.: Transport Phenomena. Wiley and Sons, Inc., New
York, 1960.
[10] Srinivasan S., Tannehill J .C., Weilmuenster K.J.: Simplified Curve Fits for the Transport
Properties of Equilibrium Air. Iowa State Univ., ISU-ERI-Ames 88405, 1987.
[11] Baldwin, B.; Lomax, H.: Thin Layer Approximation and Algebraic Model for Separated Thrbulent Flows. AIAA-paper, No. 78-257, 1978.
[12] Vandromme D.: Thrbulence Modelling and Implementation in Navier-Stokes Solvers. VKI
Lecture Series LS 1989-06, von Karman Institut for Fluid Dynamics, Rhode-Saint-Genese,
1989.
[13] Gokcen, T., Mac Cormack, R.W., Chapman, D.R.: Computational fluid dynamics near the
continuum limit. AIAA paper, No. 87-1115, 1987.
[14] Dutt. P.: Stable Boundary Conditions and Difference Schemes for Navier-Stokes Equations.
ICASE report No. 85-37, 1985.
[15] Nordstrom, J.: Energy Absorbing Boundary Conditions for the Navier-Stokes Equations. In:
Lecture Notes in Physics, vol 264, Springer Verlag, 1986.
[16] Lax, P.D., Richtmeyer, R.D., Comm. on Pure and Appl. Math., vol 9, 1956.
[17] Jameson, A., Baker, T.J.: Euler Calculations for a Complete Aircraft. Lecture Notes in Physics,
vol. 264, Springer-Verlag Berlin, 1986.
[18] Rossow C.: Comparison of Cell Centered and Cell Vertex Finite Volume Schemes. In: Deville,
M. (Ed.), Notes on Numer. Fluid Mech., vol 20, pp. 327-334, Vieweg Verlag 1988.
[19] Dortmann, K.: Computation of Viscous Unsteady Compressible Flow about Airfoils. In: Proc.
of the ICllNMFD, Lecture Notes in Physics, Springer Verlag, 1989.

68

Computational Methods in Hypersonic Aerodynamics

[20] Lohner, R., Morgan, K., Zienkiewicz: The Solution of Non-Linear Hyperbolic Equation Systems by the Finite Element Method. Int. J. of Numer. Methods in Fluids, vol. 4, pp. 1043-1063,
1984.
[21] Mavriplis D. J., Jameson A.: Multigrid Solution of the Navier-Stokes Equation on Triangular
Meshes. AIAA Journal, vol. 28, pp. 1415-1425, 1990.
[22] van Leer, B.: Towards the Ultimate Conservative Difference Scheme. A second-order sequel to
Godunov's method. J. Compo Phys. vo1.32, pp.101-136, 1979.
[23] Harten, A.: High Resolution Schemes for Hyperbolic Conservation Laws, J. Compo Phys., vol.
49, pp. 357-393, 1983.
[24] Yee H. C.: A Class of High-Resolution Explicit and Implicit Shock-Capturing Methods. In:
VKI Lecture Series 1989-04, Rhode-Saint-Genese, 1989.
also: Yee H. C.: Upwind and Symmetric Shock-Capturing Schemes. NASA TM-89464, 1987.
[25] Godunov, S.K.: Finite-Difference Method for Computation of Discontinuous Solutions. (in
Russian), Math. Sbornik vol.47, pp 271, 1959.
[26] Collela, P., Glaz, H.M.: Efficient Solution Algorithms for the Rieman Problem for Real Gases.
J. Compo Phys., vol 59,pp 264, 1985.
[27] van Leer, B.: Flux-Vector Splitting for the Euler Equations. Lecture Notes in Physics vol. 170,
pp. 507-512, 1982.
[28] Roe, P.L.: Approximate Riemann Solvers, Parameter Vectors and Difference Schemes. J. Compo
Phys., vol. 22, pp. 357, 1981.
[29] Osher, S., Chakravarthy, S.: High Resolution Schemes and Entropy Condition. SIAM J. Num.
Anal., vol 21 pp. 955-984, 1984.
[30] Steger, J .L., Warming, R.F.: Flux-Vector Splitting of the Inviscid Gas Dynamic Equations
with Applications to Finite-Difference Methods. J. Compo Phys., vol 40, pp 263-293, 1981.
[31] van Leer, B., Thomas, J. L., Roe, P. L., Newsome, R. W.: A Comparison of Numerical Flux
Formulas for the Euler and Navier-Stokes Equations. AIAA paper 87-1104 CP 1987.
[32] Eberle A., Schmatz, M., Bissinger, N.: Generalized Fluxvectors for Hypersonic Shock Capturing. AIAA paper No. 90-0390, 1990.
[33] Thomas, J.L., Walters, R.W.: Upwind Relaxation Algorithms for the Navier-Stokes Equations.
AIAA-paper No. 85-1501, 1985.
[34] Schroder, W., Hanel, D.: An Unfactored Implicit Scheme with Multigrid Acceleration for the
Solution of the Navier-Stokes Equations. Compo &. Fluids, vol 15, pp. 313-336, 1987.
[35] Schwane, R., Hanel, D.: An Implicit Flux-Vector Splitting Scheme for the Computation of
Viscous Hypersonic Flow. AIAA-paper No. 89-0274, 1989.
[36] van Leer B.: Flux Vector Splitting for the 1990's. CFD Symp. on Aeropropulsion, NASA Lewis
Res. Center, Cleveland, 1990.
[37] Seider G., Hanel D.: Numerical Influence of Upwind TVD Schemes on Transonic Airfoil Drag
Prediction. AIAA paper 91-0636, 29th Aerospace Sciences Meeting, Reno (USA), Jan. 1991,
1991.
[38] Jameson, A., Schmidt, W., Turkel, E.: Numerical Solution of the Euler Equations by FiniteVolume Metaods Using Runge-Kutta Time-Stepping Schemes. AIAA-paper 81-1959, 1981.

Computational Methods in Hypersonic Aerodynamics 69


[39] Eliason, P., Rizzi, A.: Hypersonic Leeside Flow Computations Using Centered Schemes for
Euler Equations. In: Notes on Numerical Fluid Mechanics, vol. 29, Vieweg Verlag, 1990.
[40] Vatsa V. N., Wedan B. W.: Developement of a Multigrid Code for 3-D Navier-Stokes Equations
and its Apllication to a Grid-Refinement Study. Computers & Fluids, vol. 18, pp. 391-403, 1990.
[41] Radespiel R., Rossow C., Swanson R.C.: Efficient Cell Vertex Multigrid Scheme for the ThreeDimensional Navier-Stokes Equations. AIAA Journal, vol. 28, pp. 1464-1472, 1990.
[42] Deconinck, H., Struijs, R., Roe, P.L.: Fluctuation Splitting for Multidimensional Convection
Problems: an Alternative to Finite Volume and Finite Element Methods. VKI Lecture Series
1990-03, 1990.
[43] Powell, K., van Leer, B.: A Genuinely Multi-Dimensional Upwind Cell-Vertex Scheme for the
Euler Equations. AIAA-paper 89-0095, 1989.
[44] Rumsey C., van Leer B., Roe P.: A Grid- Independent Approximate Rieman Solver with
Applications to the Euler and Navier-Stokes Equations. AIAA paper 91-0239, 29th Aerospace
Sciences Meeting, Reno (USA), Jan. 1991, 1991.
[45] van Albada, G. D., van Leer, B., Roberts, W. W.: A Comparative Study of Computational
Methods in Cosmic Gas Dynamics. Astron. Astrophys. vol. 108, pp.76-84, 1982.
[46] Roe, P.L., Baines, M.J .:Algorithms for Advection and Shock Problems. Proc. of 4th GAMM
Conf. on Num. Meth. in Fluid Mech., Notes on Num. Fluid Mech., vol 5, Vieweg Verlag 1982.
[47] Sweby, P.K.: High Resolution Schemes Using the Flux Limiter for Hyperbolic Conservation
Laws, SIAM J. Numer. Analyt., vol. 21, pp.995-1011, 1984.
[48] Grossman B., Walters R. W.: Analysis of Flux-Split Algorithms for Euler's Equations with
Real Gases. AlA A Journal, vol. 27, pp. 524-531, 1989.
[49] Montagne J.L., Yee H.C., Vinokur M.: Comparative Study of High- Resolution Shock- Capturing Schemes for a Real Gas. Notes on Numerical Fluid Mechanics, vol. 20, Vieweg Verlag,
Braunschweig/ Wiesbaden, 1988.
[50] Kroll N., Gaitonde D., Aftosmis: A Sytematic Comperative Study of Several High Resolution
Schemes for Complex Problems in High Speed Flows. AlA A paper 91-0636, 29th Aerospace
Sciences Meeting, Reno, Jan. 7-10, 1991.
[51] Hanel, D.: On the Accuracy of the Upwind Schemes in Solutions of the N avier-Stokes Equations, AIAA paper No. 87-1105 CP 1987.
[52] Hanel D.: Effects of Numerical Dissipation in Solutions of the Navier-Stokes Equations. Proc.
of Third Intern. Conf. on Hyperbolic Problems. Uppsala, Sweden, June 11-15, 1990.
[53] Mac Cormack, .W.: The Effect of Viscosity in Hypervelocity Impact Crate ring. AIAA paper
69-354, 1969.
[54] Shang, J.S., Scherr, S.J.: Numerical Simulation of the Flowfield around a Complete Aircraft.
AGARD Conf. Proc. 412, 1986.
[55] Beam, R. M., Warming, R.F.: An implicit factored scheme for the compressible Navier-Stokes
equations. AIAA J. vol. 16, pp. 393-402, 1978.
[56] Rieger, H., Jameson, A.: Solution of steady three-dimensional compressible Euler and NavierStokes equations by an implicit LU scheme. AIAA-paper 88-0619 1988.
[57] Pulliam T. H., Steger J. L.: Implicit Finite-Difference Simulations of Three-Dimensional Flow.
AIAA Journal, vol. 18, pp. 159-167, 1980.

70 Computational Methods in Hypersonic Aerodynamics


[58) Chausee D. S., Pulliam T. H.: Two-Dimensional Inlet Simulation using a Diagonal Implicit
Algorithm. AIAA Journal, vol. 19, pp. 153-159, 1981.
[59) Briley W. R., McDonald H.: On the Structure and Use of Linearized Block Implicit Schemes.
J. of Comput. Physics, vol. 34, pp. 34-72, 1980.
[60) Bardina, J., Lombard, C.K.:Three Dimensional Hypersonic Flow Simulations with the CSCM
Upwind Navier-Stokes Method. AlA A paper 87-1114 CP, 1987.
[61) Yokota J. W.: Diagonally Inverted Lower-Upper Factored Implicit Multigrid Scheme for the
Three-Dimensional Navier-Sokes Equations. AlA A Journal, vol. 28, pp. 1642-1649, 1990.
[62) Shuen J. S.: Numerical Study of Chemically Reacting Flows Using a Lower-Upper Symmetric
Successive Overrelaxation Scheme. AIAA Journal, vol. 27, pp. 1752-1760, 1989.
[63) Candler G. V., MacCormack R. W.: Hypersonic Flow past 3-D Configurations. AIAA paper
AIAA-87-0480, 1988.
[64) Lacor C., Hirsch C.: 3-D Computations of Complex Flow Systems. Proc. of 16th ICAS Congr.,
1988.
[65) Riedelbauch, S., Brenner, G.: Numerical Simulation of Laminar Hypersonic Flow past Blunt
Bodies Including High Temperature Effects. AIAA paper No. 90-1492, 1990.
[66) Schroder W. & Hartmann G.: Robust Computation of 3-D Viscous Hypersonic Flow Problems,
Proc. of the 9th GAMM Conf. on Num. Meth. in Fluid Mech., Notes on numer. Fluid Mech.,
Vieweg Verlag, 1991.
[67) Jameson, A.: Solution of the Euler Equations for Two-Dimensional Transonic Flow by a Multigrid Method. Appl. Math. and Comput., vol. 13, pp 327-355, 1983.
[68) Chima R. V., Yokota J. W.: Numerical Analysis of Three-Dimensional Viscous Internal Flows.
AIAA J., vol 28, pp. 798-806, 1990.
[69) Meinke M., Hanel D.: Time Accurate Multigrid Solutions of the Navier-Stokes Equations.
Intern. Series of Numerical Mathematics, vol. 98, pp. 289-300, Birkhauser, Basel, 1991.
[70) Lohner, R., Baum, J.D.: Numerical Simulation of Shock Interaction Using a new Adaptive
H-Refinement Scheme on Unstructured Grids. AIAA paper No. 90-0700, 1990.
[71) Baker, T.J., Jameson, A.: Improvements to the Aircraft Euler Method, AlA A paper No.
87-0452, 1987.
[72) Peraire J., Formaggio L., Morgan K., Zienkiewicz: Finite Element Euler Computations in
Three Dimensions. AIAA paper No. 88-0032, 1988.
[73) Mavriplis D.: Turbulent Flow Calculations using Unstructured and Adaptive Meshes. ICASE
Rep. No. 90-61, 1990.
[74) Nakahashi N.: FDM- FEM Zonal Approach for Viscous Flow Computations over Multiple
Bodies. AIAA paper No. 87-0604, 1987.
[75) Wheatherill N .P.: A Strategy for the Use of Hybrid Sructured- Unstructured Meshes in CFD.
In: Numerical Methods for Fluid Dynamics (3), Ed. Morton K. W., Baines M.J., Oxford
University Press, 1988.
[76] Hassan, 0., Morgan, K., Peraire, J.: An Adaptive Implicit/Explicit Finite Element Scheme for
Compressible Viscous High Speed Flows. AIAA paper No. 89-0363, 1989.
[77) Prabhu, P.K., Stewart, J .R., Thareja, R.R.: A Navier-Stokes Solver for High Speed Equilibrium
Flows and Application to Blunt Bodies. AIAA paper 89-0668, 1989.

Computational Methods in Hypersonic Aerodynamics 71


[78] Lohner R., Morgan K., Peraire j., Vahdati M.: Finite Element Flux-Corrected Transport for
the Euler and Navier-Stokes Equations. Int. J. for Numer. Meth. in Fluid Mech., vol. 7, pp.
1093-1109, 1987.
[79] Periaux, J.: Finite Element Simulations of Three-Dimensional Hypersonic Reacting Flows
around Hermes. Contrib. on Second Joint Europe-US Short Course on Hypersonics. Org. by
GAMNI-SMAI, Univers. of Texas at Austin, and U.S. Air Force Academy, Colorado Springs,
Jan. 1989,1989.
[80] Mallet M.: Adapted Finite Element Methods for Hypersonic Reentry Problems. In: Third
Joint Europe/US Short Course in Hypersonics, RWTH Aachen, Oct. 1990, 1990.
[81] Venkatakrishnan V., Barth T.J.: Application of Direct Solvers to Unstructured Meshes for the
Euler and Navier-Stokes Equations Using Upwind Schemes. AIAA paper No. 89-0364, 1989.
[82] Desideri J .A., Glinsky N., Hettena E.: Hypersonic Reactive Flow Computations. Computer &
Fluids, vol. 18, pp. 151-182, 1990.
[83] Rostand P., Stouflett B.: TVD Schemes to Compute Compressible Viscous Flows on Unstructured Grids. In: Notes on Numerical Fluid Mechanics, vol. 24, pp. 510-520, Vieweg Verlag,
Braunschweig/ Wiesbaden, 1989.
[84] Brandt, A.: Multi-Level Adaptive Solutions to Boundary-Value Problems. Mathematics of
Computation, vol 31, No. 138, pp. 333-390, 1977.
[85] Brandt, A.: Guide to Multigrid Development. In: Lecture Notes in Mathematics vol 960., pp.
220-312, Springer Verlag Berlin, 1981.
[86] Chima, R. V., Johnson, G. M.: Efficient Solution of the Euler and Navier-Stokes Equations
with a Vectorized Multiple-Grid Algorithm, AIAA Journal, vol. 23, No.1, pp. 23-32. 1985.
[87] Ni, R. H.: A Multiple Grid Scheme for Solving the Euler Equations, AIAA Journal, vol. 20,
No. 11, pp. 1565-1571 1982.
[88] Hemker, P. W., Spekreijse, S. P.: Multiple Grid and Oshers Scheme for the Efficient Solution
of the Steady Euler Equations. Appl. Num. Math., vol. 2, pp475-493, 1986.
[89] Mulder, W. A.: Multigrid Relaxation for the Euler Equations. J. Compo Phys., vol 60, pp.235252,1985.
[90] Shaw, G., Wesseling, P.: Multigrid Method for the Compressible Navier-Stokes Equations.
Rep. of the Dep. of Math. and Inf., Nr. 86-13, Univ. Delft, 1986.
[91] Anderson, W. K., Thomas, J.L., Rumsey, C. L.: Extension and Application of Flux-Vector
Splitting to Calculation on Dynamic Meshes, AlA A paper 87-1152 CP 1987.
[92] Hemker, P.W., Koren, B.: Multigrid Defect Correction and Upwind Schemes for the Steady
Navier-Stokes Equations. In: Proc. of Third Conf. on Numerical Methods for Fluid Mech.,
(Morton K. W., Baines M. J., Editors), Oxford, 1988.
[93] Radespiel, R., Swanson, R.C.: An Investigation of Cell Centered and Cell Vertex Multigrid
Schemes for the Navier-Stokes Equations. AIAA paper 89-0548, 1989.
[94] Siciari, M.J., Del Guidice, P., Jameson, A.: A Multigrid Finite Volume Method for Solving the
Euler and Navier-Stokes Equations for High Speed Flow. AIAA paper 89-0283 1989.
[95] Abid R., Vats a V. N.: Prediction of Separated Tronsonic Wing Flows with Nonequilibrium
Algebraic Turbulence Model. AIAA Journal, vol. 28, pp. 1426-1431, 1990.

72

Computational Methods in Hypersonic Aerodynamics

[96] Gerolymos G. A.: Implicit Multiple Grid Solution of the Compressible Navier-Stokes Equatios
Using k- Turbulence Closure. AIAA Journal, vol. 28, pp. 1707-1714, 1990.
[97] Meinke M., Hanel D.: Application of the Multigrid Method in Solutions of the Compressible
Navier-Stokes Equations. Proc. of the Fourth Copper Mountain Conf. on Multigrid Methods.
(Mandel J. et al, Editors) SIAM, Philadelphia, 1989.
[98] Harten A.: On a Class of High Resolution Total-Variation-Stable Finite-Difference Schemes.
SIAM J. Num. Anal., vol. 21, pp 1-23" 1984.
[99] R. Schwane, D. Hanel: An Upwind Relaxation Method for Hypersonic Flows. Computer &
Fluids, vol. 19, 1991.
[100] Workshop on Hypersonic Flows for Reentry Problems. Organized by INRIA and GAMNISMAI, Jan. 1990, Antibes, France, 1990.
[101] Schwane, R., Hanel, D.: Computation of Viscous Supersonic Flow around Blunt Bodies. 7th
GAMM Conf. on Num. Math. in Fluid Mech., Louvain-La Neuve, 1987.
[102] Riedelbauch, S., Miiller, B., Rues, D.: 3D Viscous Hypersonic Flow over Blunt Bodies. DVLRFB 89-09, 1989.
[103] Schmatz, M.: Hypersonic Three-dimensional Navier-Stokes Calculations for Equilibrium Gas.
AIAA paper No. 89-2183, 1989.
[104] Miiller, B.: Development of an Upwind Relaxation Method to Solve the 3D Euler and NavierStokes Equations for Hypersonic Flows. GAMM Wissenschaftliche Jahrestagung, Hannover,
1990.
[105] Riedelbauch, S.,Miiller, B., Kordulla, W.: Semi-Implicit Finite Difference Simulation of Laminar
Flow over Blunt Bodies. Proc. of 11th ICNMFD, Williamsburg, 1988.
[106] Prabhu, D.K., Tannehill, J.C., Marvin, J.G.: A New PNS Code for Chemical Nonequilibrium
Flows. AIAA-J .,vol 26, No.7, pp. 808-815, 1988.

---

----------------

Computational Methods in Hypersonic Aerodynamics 73

100~------------------

______________~

,/,;;.;;~:~> .
.,.,.-

_ --

------,.----

C'"

~~ = 0
o

" 0

"o--:;-';'COOoMO

10" ,,,,,,,, .. ,, "

No-slip

Moo = 20,

0,6 - 0,8 Experiment

0,7

Generalized slip

~~ = 0

Maxwell slip

0,7

la'

Monte Carlo

~~ = 0

10'

Generalized slip + T. - Tw
10'
Re oo

10'

10'

Fig. 1: Skin friction drag of a flat plate versus Reynolds number.


Influence of the slip conditions in near-continuum flow [13] .

Oi,j , i'i,j

node -centered

Oi,j, ri,j

cell-vertex

x Oi,j , ri,j

cell- centered

Fig. 2: Different arrangements of the control volume.

74 Computational Methods in Hypersonic Aerodynamics

0.08

(1)
(2)
(3)
(I.)

-3

-I.

(5)

')(. = 1
')(. = 113
')(.=0
')(. =-1

Ijl = 1
Ijl = 1
Ijl = 1
Ijl = 1
Ijl=O

-5

Fig. 3: Relative error between numerical and exact solution of the Burgers equation.
Influence of different discretizations, using the MUSCL approach.

2
U-U.,x.

u.,x

0.08
-1

-2
(1 )

'l<.=

a }

(2) 'l<.= 113


(3) 'l<.=1

-3

0) IP = 1

b) IP = IPv.AlI!Al:lA

-I.

Fig. 4: Relative error between numerical and exact solution of the Burgers equation.
Influence of the flux limiter, using the MUSCL approach.
- - no limiter (cp = 1)
- - - - with limiter (cp =

-- -

CPv.Albada)

------------

Computational Methods in Hypersonic Aerodynamics 75

u:
0

U1

-Cp

0
C

U1

0
I

Experiment
Cook et aI., 1979

FDS solution
minmod-limiter with w = 2
.------- minmod-limitcr with w = I

0
~

I
U1
~

0.0 0.2 0.4 0.6 0.8

1.0

x/c
Fig. 5: Computed pressure distribution for transonic, turbulent flow around a
RAE 2822 airfoil [37]. (Re = 6.5 x 106 , Ma oo = .73, 0: = 2.79).
Influence of different min-mod flux limiters:

k'l}

'P ,. = minmod{2wr.
l+r'l+r'

= 2:
W = 1:

--W
- - - -

Cl
Cl

to.

= .823 , Cd = .0187
= .820 , Cd = .0216

05
-Blasius

Cf

o Van Leer
6.

0.()I.

Vee/Harten

17<33
Rea:> :. lOS
Ma~

0.03

= .5

0.02

0,01

0.5

X/L _

1.0

Fig. 6: Solution of the 2-D Navier-Stokes equations for laminar flow over a flat plate.
Influence of different upwind formulations on the values of skin friction:
flux-vector splitting, van Leer [27],
flux-difference splitting, Vee [24],

O(~2)

O(~2)

76 Computational Methods in Hypersonic Aerodynamics

Flux-Vector-SplLtting

0
6.

Blasius
Van Leer
present

0,03
o

0.02

0.01

0.1

0.2 0.3

0.4 0.5 0.5 0.7

0.8 0.9

1.0

x/L-

Fig. 7: Solution of the 2-D Navier-Stokes equations for laminar flow over a flat plate.
Influence of different formulations for flux-vector splitting on the values of
skin friction:
original flux-vector splitting by van Leer [27]
modified splitting by Hanel, Schwane [35]

a)

Fig. 8: Navier-Stokes solution for hypersonic flow around a double-ellipsoidal body.


Vectors of the local skin friction on the body surface.
(Ma oo = 8.15, Re oo = 106 , a = 30,60.000 grid points).
a) van Leer flux-vector splitting [27]

b) splitting by Hanel, Schwane [35]

Computational Methods in Hypersonic Aerodynamics 77

Ma.,. = 2.91.
Re.,. = 2.2)(105

1,1

Pr

=.72

---- present, V. LEER fluxes


- - present, modif. fluxes
VIVIANO et aI.
o

1,0

0.9

2._

Fig. 9: Wall temperature distribution related to the inflow stagnation temperature,


over the arc length in the mid plane of a hemisphere-cylinder body.
van Leer flux-vector splitting [27]
modified split energy flux, Hanel [51J
central scheme by Viviand et al.

Fig. 10: Navier-Stokes solution for the hypersonic, laminar flow during the stage separation of a two-stage space vehicle. (M a oo = 6., Re/m = 4.2 x 10 5 , 0: = 0).
a) The two surface grids
b) Lines of constant density in the symmetry plane

78 Computational Methods in Hypersonic Aerodynamics

Fig. 11: Computation of the inviscid and viscous flow around a space shuttle, [56].
(Ma = 8,Re/m = 106 ,0: = 30).
Isomach lines in the cross section x = 12.7m for viscous (left) and inviscid
flow (right).

t.Orrr-------------,

CfX 10

--NS
- ----- PNS
- - - RNS

HOLDEN AND
MOSELLE

0.5

OL-_ _ _ _ _

______
XlL

Fig. 12: Comparison of the skin friction over a compression ramp calculated
using the NS, PNS, and RNS equations. (Results taken from [6]).

Computational Methods in Hypersonic Aerodynamics 79

b)

Fig. 13: Computation of hypersonic flow around a two-stage vehicle with a space
marching method [4]. (M a = 6.8, Re oo = 8 x 106 ,0: = 80 )
a) Isobars of inviscid flow on the upper surface
b) Isobars of inviscid flow in cross-section z

= 4.2

c) Isobars of viscous flow in cross-section z = 4.15

Chapter 3:
Numerical Simulation of Hypersonic Flows
J. S. Shang
Wright Laboratory, Wright-Patterson Air Force Base,
Ohio 45433-6553, USA
ABSTRACT
Numerical simulation of aerodynamic and chemical kinetic phenomena encountered
in the hypersonic flow regime is outlined. In the continuum flow region, the current
numerical methods focus on solving the Navier-stokes equations in conjunction with
chemical physics equations, including the non-equilibrium chemical reactions and internal degrees of excitation. In the rarefied gas domain, the state-of-the-art progress
using particle simulation which does not rely on continuum concepts, such as the Direct Simulation Monte Carlo (DSMC) method, has gained confidence but is still in
the embryonic state, thus will only be briefly delineated. Similarly, the added mechanism of energy transport process in radiation is presented in the same line of thought.
UNIQUE FEATURES OF HYPERSONIC FLOWS
Rarefied Gasdynamics
In practical applications, the flight envelope of hypersonic vehicles encompasses the
entire domain of gas dynamics from free molecule to continuum regime. In these two
relatively well defined realms, the interaction of a flight vehicle and its surrounding
transits from an intermolecular collisions dominated environment to a circumstance,
in which particle collisions with the vehicle is critical. The Knudsen number, defined
by the ratio of the mean free path and a pertaining length scale of the flowfield, is
a measure of the degree of rarefaction [1,2]. The governing equations of these two
regions are the Navier-Stokes equations and the Boltzmann equation, respectively.
The two systems of equations are closely related through the kinetic theory. The
hierarchy of the pertaining governing equations in aerodynamic applications is given
in Figure 1. Although the shock wave structure usually dominates the hypersonic
flowfield, the interaction with a viscous layer is intensified as the Mach number increases. Separate treatment of the invisicid and viscous phenomena, as if they were
independent from each other, becomes increasingly inaccurate. Therefore, the inviscid
and the boundary-layer approximations have a rather limited range of applicability in
hypersonic flows. To define hypersonic flows by the ratio of the flow velocity relative

82 Computational Methods in Hypersonic Aerodynamics


to the vehicle and the ambient speed of sound is an over simplification. The characteristics of hypersonic flows vary gradually with the geometry of the vehicle, the
nature of surrounding atmosphere, and the flight velocity. Nevertheless, hypersonic
flows are generally characterized by a relatively high Mach number of the oncoming
stream.
In the continuum regime, hypersonic flows are featured by a highly compact domain of dependence within strong shock envelopes and surrounded by a medium with
high energy content. In addition to the collision mechanism, radiation contributes
to the energy transport process. Critical hypersonic phenomena can be accurately
described only by combining chemical physics with aerodynamics. The occurrence
of high temperature physical phenomena may be concentrated in a few local regions,
but their impact to the development of the entire flowfield structure is not completely
understood; particularly, when a large amount of energy either absorbed or released
by the nonequilibrium chemical reactions may control the development of the entire
flow, Figure 2. The coupling of these two distinct scientific disciplines in hypersonic
analyses is inclusive. However, unique features in hypersonic flow are traditionally
still separated into the aerodynamic and chemical-physics aspects.
Classic Hypersonic Theories
Although the inviscid small-disturbance theory [3J has a vanishing valid range of
application in hypersonics, the order of error analysis is still most instructive. For a
slender body, the disturbed streamwise velocity component is an order of magnitude
lower than the cross flow component. The observation reveals that the hypersonic
motion of a slender body through a gas produces mostly transverse displacement of
the gas and becomes the foundation of the equivalence principle [4J or the Law of
plane section [5J. In essence, for a steady, three-dimensional hypersonic gas motion
around a slender body, the lengthwise body axis can be transformed into a timelike axis. In other words, the steady inviscid hypersonic flowfield around a slender
body is identical to that of an unsteady motion with one less spatial dimension.
The approximate inviscid equations and boundary conditions reduce to that of the
unsteady gas motion in the cross-flow plane section. From this frame work, the selfsimilar or blast wave solution was developed independently by G.1. Taylor [6J and 1.1.
Sedov [7J. These illustrious efforts highlight the dynamic and inertial force dominant
nature of hypersonic flows.
As the Mach number increases, even for a slender body, the interaction between
inviscid and viscous flowfields can no longer be ignored. In the hypersonic domain, a
thick boundary layer superimposing on the body leads to the relatively large, outward
streamline deflection. The interaction of the effective body shape with the Mach waves
in the inviscid field has been identified by Hays and Probstein [2J as the pressure
interaction or the Mach-wave interaction. The induced pressure on a flat plate at zero
incidence, for either the strong or the weak interaction, is proportional to a parameter
X = VCM3/~ where C is the Chapman-Rubesin constant, M is the Mach number
and Re is the Reynolds number.
The pressure induced by viscous-inviscid interaction can easily overwhelm the
original flow structure at low supersonic speed. Hays and Probstein have shown that,
for strong pressure interaction, the boundary-layer growth rate on a plate changes

Computational Methods in Hypersonic Aerodynamics 83


from a half to the three-quarter power of the running length. Under hypersonic
conditions, the significance of vorticity interaction is also pointed out by Ferri and
Libby [2]. The effect of vorticity interaction is to increase the skin friction and heat
transfer. This phenomenon is important for a blunted slender body, especially at very
low values of the Reynolds number [2]. Under this flow condition, the thick boundarylayer on the body surface is affected not only by the external velocity profile but also
by the vorticity distribution induced by the highly curved bow shock wave structure.
Some of the hypersonic interacting fluid dynamics phenomena can be analysed by the
interacting boundary-layer theory [8], but the Navier-Stokes equations definitely have
a much broader range of validity. For example, the transverse curvature effect incurred
by the scaling of the boundary-layer approximation needs no special treatment by
solving the Navier-Stokes equations. In addition, even at a Knudsen number up to
the order of unity, accurate solutions of skin friction and heat transfer on a simple
shape have been recovered through the Navier-Stokes equations with modified surface
boundary conditions [9].
The features of continuum hypersonic flowfield around a blunt body emerge even
at low supersonic Mach numbers. A mixed local subsonic, transonic, and supersonic
domain, including the stagnation point is bounded by the bow shock wave and the
body. The stand-off distance of the detached shock wave is uniquely determined by
satisfying the continuity requirement, and is linearly proportional to the density ratio
across the shock. In general, the shock surface is smoother than the body surface
and the shock shape is relatively insensitive to small changes of the body surface.
The curved bow shock also generates a continuously varying entropy field and renders the flow within the shock layer to be highly rotational. Even though there are
distinguishable structural differences between the boundary layer and the inviscid region of relatively cool high density flow within the shock layer, the entire flowfield is
dominated by viscous effects [2,10]. However, the high pressure gradient within the
layer is mostly generated by the inviscid mechanism. From an eigenvalue analysis of
the inviscid equations, some of these eigenvalues degenerate to zero in a short session
from each other. As it will become clear later, this peculiarity will create a numerical
resolution issue for numerical simulations.
Chemical Physics
The most outstanding property of hypersonic flows lies in the chemical physics of
the flow medium encountered in the high temperature environment. The high temperature and pressure condition is the direct consequence of the energy conversion
process brought about by extremely strong shock waves. In this confined region, all
temperature induced molecular/atomic excitations of the flow medium are governed
by collision and radiation mechanisms. Equilibrium states of the system are rarely
achievable for a wide range of speeds and altitudes of hypersonic flights. Departure
from thermodynamic and chemical equilibrium is known to have significant effects on
the flowfield structure and the propagation of disturbances [11,12]. The thermodynamic properties of an air mixture differs considerably from that of a perfect gas. At
a temperature of around 1500 K, the excitation of the vibrational degree of freedom
of the diatomic molecules, Oxygen and Nitrogen, becomes significant. As the temperature is increased, Nitric Oxide is formed by chemical reaction and followed by

84 Computational Methods in Hypersonic Aerodynamics


the dissociation of Oxygen, Nitric Oxide and Nitrogen. Only when the temperature
is elevated beyond 9000 K, does the ionization of air components occur. In addition
to the influence of the change in air composition with respect to temperature, the
molecular transport processes are affected by the species concentration, temperature,
and pressure [13,14]. In a multi-component reacting system, anyone species moves
at a different velocity than the mass-averaged value, a flux of mass and energy is
carried across the control surface by the diffusion process. The diffusion phenomenon
substantially alters the heat transfer characteristics in hypersonic flow.
In linking the details of atomic and molecular structure to the thermodynamic
behavior of a gas mixture, statistically, the dynamics of a real gas is considered to
consist of translational, rotational, vibrational, electronic, and nuclear internal degrees
of freedom [11,14]. For the translational motion, the majority of quantum mechanical
energy levels are closely spaced, relative to the datum level. Thus, the energy states
are nearly continuously distributed. Since only a few intermolecular collisions are
required to achieve equilibrium in the translational and rotational modes, these two
internal energy states are generally considered to be fully excited and in thermodynamic equilibrium.
The separation of the internal structure beyond translational excitation becomes
increasingly uncertain. For example, the centrifugal field of rotation affects vibration,
and the vibration motion changes the moment of inertia of rotational motion [11]. Vibrational excitation and dissociation is also closely coupled [15,16,17], because energy
is stored in the vibrations and the rate of dissociation is dependent on the vibrational state of molecules. There are two possible mechanisms to transit a molecule
from vibrational states into the dissociated state of free atoms. One is the climbingthe-vibrational-ladder theory, the other is the theory of direct transition by violent
collision with heavy particles. Regardless of the uncertainty in these theories and
known discrepancy from experimental evidence [11 ,14,17], it is recognized that the
vibrational relaxation precedes the onset of dissociation, and the rate of dissociation
is strongly dependent upon the vibrational state of the molecule. The least understood internal degree of freedom of a real gas is ionization, because free charges are
produced by the long range electrostatic Coulomb field [14]. Ionization or nuclear excitation normally has little effect on the energy balance of a flowfield, but the presence
of charged particles can be important for forced diffusion and radiation phenomena.
The avalanche of ionization phenomenon has been described by Park et at. [18]. The
basic event can be described by the associative ionization process, the transfer of
electric charge by the charge-exchange reaction and the electron impact to neutral
atomic species. Since the next electronic energy level lies far above the ground state,
the contribution from ionization to the thermodynamic properties of gas is negligible
below 10,000K [14].
In spite of the complex, interconnected internal degrees of freedom of a molecule,
the energy state can be described by the factorization property of partition functions. For weakly interacting particles of permissible internal quantum states, only
the translational degree of freedom yields a net change of momentum across the control volume. As a result, only the translational motion contributes to the pressure. In
a full equilibrium situation, the pressure is the hydrostatic pressure, a scalar quantity.
However, the pressure is a second-rank tensor in a non-equilibrium state [14]. There-

Computational Methods in Hypersonic Aerodynamics 85


fore, non-equilibrium effects in hypersonic flows are not necessarily limited to the
introduction of additional length (time)-scale dependence into the flowfield or change
the dynamics of flow by changing gas temperature.
Another unique feature of hypersonic flow accentuated by non-equilibrium state
is the surface catalytic effect. The basic phenomenon is generated by the release of
heat of formation on a surface by the recombination of atoms or radicals of chemical reaction that are conveyed by the diffussion process [19,20]. The recombination
process is controlled by the kinetics of chemical reactions in which chain reactions
may begin, as well as end on surfaces by five different mechanisms [19]. All the
previously mentioned transport phenomena, including radiation, arise as a result of
either thermodynamical or chemical non-equilibrium conditions. Unfortunately, all
these nonequilibrium phenomena are studied in terms of pertubations from the state
of equilibrium. In a non-equilibrium state, all irreversible processes are defined by
local or quasi equilibrium state variables. In general, these instantaneous values must
be regarded as purely fictitious and non-unique [11,14]. The accuracy of numerical
simulations based on these physical models at a state of significant departure from
equilibrium is highly uncertain.
Shear Layer Instability
Finally, an often overlooked, unique stability characteristic of the hypersonic boundarylayer also exists. From hypersonic experiments [21,22]' the first mode, or TollmeinSchlicting mode, carries over into supersonic and hypersonic regimes. The second
mode, which is the additional acoustic mode, becomes increasingly dominant and is
the more unstable of the two principal modes of instability in a hypersonic stream.
If the initial disturbance spectrum is nearly infinitesimal and random, with no discrete frequency peaks, the initial instability will occur as two-dimensional TollmeinSchlicting waves, traveling in the mean flow direction. The second-mode waves soon
become detectable by contained transverse flow properties. A shear layer in the unstable region has a strong ability to amplify any slight, three-dimensional small disturbance that may appear in any nature-disturbance spectrum. From experiments,
the second-mode waves exhibit significant amplification as the Mach number increases
beyond the value of four [21]. The largest disturbance in a hypersonic boundary layer
is near the outer edge of this layer, which coincides with the generalized inflection
point of the boundary-layer profiles [22]. Although there is no established theory of
laminar-turbulent transition, transition is preceded by the selective amplification of
small disturbances. Initial disturbances that may trigger the transition appear in a
composite spectrum of intensities and frequencies that have been attributed to numerous parameters and possible external sources. Trends of hypersonic transition have
exhibited some contradictions between theory and stability experiments [22]. The
added effects of chemical physics to transition phenomena are still open for future
research.
GOVERNING EQUATIONS
Computing technology contributes to two areas of hypersonic research [23]. First, it
provides an added dimension to simulate physics that eludes the current physical mea-

86 Computational Methods in Hypersonic Aerodynamics


suring capabilities in a high temperature and a controlled environment, particularly
for flowfields dominated by fine scale structures and in transient states. Second, it
greatly accelerates the data acquisition and reduction process. In order to understand
and interpret the numerical results of theoretical modeling, many intermediate steps
in data manipulating and managing are necessary. These tasks can only be accomplished with the aid of computer systems. Computational aerodynamics in hypersonic
regime is not merely a modest extension of conventional computational fluid dynamics, but rather an integrated scientific discipline including complex chemical physics
as they have been discussed earlier. The added description of chemical physics of gas
at high temperature must be incorporated into the equations of motion to form a
basis of interdisciplinary computational fluid dynamics.
Microscopic and Macroscopic Formulations
If the nature of the microscopic process were known then the motion of the gas could
be completely described by the detailed collision process. This is the basic approach
of the kinetic theory of gas, which follows the dynamic trajectories of individual
particles from known initial conditions. The fundamental equation of kinetic theory
of gas is the Boltzmann equation [11,12], and is an integro-differential system. This
equation describes the rate of change with respect to position and time of the velocity
distribution function based on the general concept of statistics. In kinetic theory,
the position and velocity coordinates are considered to be independent. Therefore,
the solution of the Boltzmann equation is probabilistic rather than deterministic. In
principle, numerical solutions of the Boltzmann equation can be obtained through
Monte Carlo methods.
For hypersonic applications, small perturbation and Euler approximations to
equations of gas motion are of limited value because of their inherent inability to
describe the pertinent physical phenomena. In the region of continuum mechanics,
the ensemble averaged Navier-Stokes equations and the reduced Navier-Stokes equations (Thin-Layer Navier-Stokes, Parabolized Navier-Stokes and Shock-layer equations) are still the mainstays. This macroscopic conservation system of equations
aided by modified boundary conditions has demonstrated useful extensions of the
continuum concept beyond its theoretical limit [9,24]. In the continuum formulation
leading to the Navier-Stokes equations, the differential system is not closed. Additional auxiliary equations are needed to describe the global behavior of the collision
processes through statistical means - the transport properties, equations of state, rate
of chemical reactions and radiation heat transfer are needed to describe all pertaining
physical phenomena in hypersonic flows.
Diffusion Phenomena and Transport Properties
In a gas mixture, the diffusion phenomenon is well known. From the kinetic theory of
diluted gases, the expression of the diffusion coefficient is not affected by the presence
of internal degree of freedom in the molecule [25]. However, in a chemically reacting system, the diffusion phenomenon can dominate the mass and energy transport
processes. The net rate of production of chemical reactions will appear as a source
term in the conservation equations for both homogeneous and heterogeneous reacting
systems. Especially for heterogeneous reactions, the chemical change takes place only

Computational Methods in Hypersonic Aerodynamics 87


in a restricted region, such as the surface of ablative materials. The rate of production will also appear in the boundary condition at the solid surface. The diffusion
phenomenon is one of the major added complications in the analysis of hypersonic
flows. In a diffusing mixture, the velocities of individual species can be significantly
different from each other. The species diffusion velocity (Vi) is defined by a relative
velocity with respect to a local mass-averaged velocity of the mixture, U [13,25].

(1)
where Pi is the density of species i. The relative diffusion velocity of a chemical species
(Ui) is defined by the difference of the mass-averaged value and the species diffusion
velocity with respect to a stationary frame of reference.
Ui

= Vi -

(2)

By definition, the sum of mass diffusion velocities of all species is identical to zero,
= O.
For a multi-component chemically reacting system, the concentrations of various species are defined as: mass concentration, Pi, mass fraction, 0i = p;/ p, molar
concentration, Xi = p;/Mi' Mi denoting molecular weight, and the mole fraction,
f3i = X;/ I:i Xi. The fractions have the following properties, I: 0i = 1 and I: f3i = 1.
The relationship between mass and mole fraction is

I: PiUi

(3)
The diffusive mass flux is generated by three mechanical driving forces and a contribution by temperature gradient [13,25]. The forced diffusion of a species can be induced
by an external force exerted on electrically charged particles, and is important only
in an ionized gas system where the electric field is present.
In general, the mass flux vector is expressed by the first-order spatial derivatives
of various physical properties and mass-averaged velocity. Although the diffusive flux
vectors were originally derived for monatomic gases at low density condition, they have
been applied to polyatomic gases with little error [25]. According to Bird, Stewart,
and Lightfoot [13], the diffusive mass fluxes are given as:
Ordinary diffusive mass flux

(4)
",here c, Rand T denote the speed of sound, the gas constant and temperature
respectively, and with P denoting pressure and S denoting entropy, G i is the Gibbs
free energy, G; = E; + PV - TS.
Pressure diffusive mass flux

(5)

88

Computational Methods in Hypersonic Aerodynamics

Forced diffusive mass flux

(6)
Thermal diffusive mass flux

(7)

= - DrlnT

PiUi

where Dij and DT are multi-component diffusion and thermal diffusion coefficients
respectively. The Ii on the right hand side of the force diffusion equation is the
external body force, a field force.
In aerodynamic application, the composition of air mixture has small differences
in molecular structure and weight between Nitrogen and Oxygen. In the dissociated
state, the weight difference between molecular and atomic species are also similar.
Therefore, the limiting case of a two-component system is frequently used in place of
a general diffusion relation. The binary mixture approximation reduces the diffusion
mass flux to the following [13]
PiUi

-PjUj

[8

-cMj2 Mj Dijj3j pRT


8j3i
Dr \lInT

(G-Mi
i

)
T,p

\l j3j - -Pj
P

1) ]

I - Ii ) + ( -Vi - Mi

\l P

(8)

If the forced and pressure diffusions together with thermal diffusion (Soret effect)
are neglected, then only the ordinary diffusion will contribute to the mass flux. For
hypersonic flow, the most important contribution to the mass flux is due to the concentration gradient. The relationship between diffusive mass flux and gradient of
molar concentration is known as Fick's first law of diffusion
P."U."

--

-cD
tJ

'\7
V

13
t

(9)

In essence, the species i moves relative to the mixture in the direction of decreasing
mole fraction of this species. The physical unit of mass diffusivity is the same as that
of the kinematic viscosity. In hypersonic applications, the transport properties of air
mixture are usually determined by a group of dimensionless parameters such as the
Lewis-Seminov, Prandtl, and Schmidt numbers [11,13,14,25]. Therefore, if anyone of
the coefficients of the transport properties can be found, the rest are determined by
an approximated value of these parameters.
In a multi-component system, the energy fluxes also are driven by the identical mechanisms of mass and momentum fluxes. In fact, there is a unique coupling
between mass and energy transport. In case the average velocity of any species differs from the mass-averaged velocity of the gas mixture, the individual species will
carry specific enthalpy, equal to Piujhi across the control surface where h denotes the
enthalpy. This additional energy flux is contributed by the inter-diffusion process.
Onsager's reciprocal relation of irreversible thermodynamic processes implies that if
temperature gives rise to the Soret effect in diffusion, the concentration gradient must
produce a heat flux. The coupling effect, known as the Dufour energy flux, provides
an additional contribution to the energy transfer. The Dufour effect is quite complex

Computational Methods in Hypersonic Aerodynamics 89


in nature and in general is usually of minor importance in most aerodynamic problems [13,25]. In current hypersonic numerical simulation, the Dufour effect of energy
flux is consistently neglected together with the Soret effect of mass diffusion. The
heat flux for a multi-component system including the radiation transfer becomes

(10)
where k is the thermal conductivity and qr denotes radiative heat transfer.
Internal Degree of Excitation of Gas Medium
From the thermodynamic point of view, the macroscopic state of a system must
be definable by a small set of state variables. The relationship between statistical
mechanics and thermodynamics is always discussed in terms of the partition function
(Z). Therefore, the partition function of a microscopic particle is fundamentally
important in relating atomic/molecular structure, which influence the energy level,
and thermodynamic behavior [11,14,25]. With k representing the Boltzmann constant,
(11)
Physically, it is a measure of the fraction of the total number of particles in the
system which possess an energy level /OJ. It describes how the energy is distributed
among particles of a system. Thus, the partition function has also been defined as the
normalizing constant for the quantum mechanical canonical distribution function [25].
/OJ is the energy level which quantum mechanics assigns to an atom/molecule on the
basis of its structure. An energy level may consist of all energy states having the same
values /OJ. Any energy level containing more than one state is said to be degenerate
and the number of states of degeneracy is denoted gi. Degeneracies are explicitly
related to the internal structure of the particles. Since large numbers of degrees
of freedom exist for a system, the accurate calculation of the partition function is
impractical. In applications, modeling of physics becomes necessary. Usually, the
coupling of internal degrees of freedom of gas medium is neglected, this assumption
leads to the factorization property of partition functions [11,14].
(12)
In accordance with commonly accepted convention, the energy of the ground level is
set to zero. At each level of internal degree of excitation, the quantity /O;f k has the
dimension of temperature, and is used to define the characteristic temperatures of the
internal degree of freedom. For air mixture, the composition changes according to
temperature encountered, and has increasing degrees of excitation as the temperature
rises. The energy of different internal degree of excitations is additive, the total
internal energy is then found by summing the translational, rotational, vibrational,
and electronic modes.
(13)
All the thermodynamic state variables can now be derived from the partition function.

90 Computational Methods in Hypersonic Aerodynamics


The pressure, specific internal energy, and entropy are given by

P= NkT (~aZ)
zav
E = NkT2 (~ aZ)
ZaT v

(14)

S = Nk {In

(~) + 1 + ~ (~~) v}

(15)

(16)

From quantum mechanics, the partition functions of translational and rotational motion are well known. The translational energy levels are very closely spaced, the
distribution of energies can be considered to be continuous. Here, h denotes Planck's
constant.
27rmkT) ~
(17)
Ztrs = ( h 2
V
The partition function of rotational degree of excitation of a diatomic gas, based on
rigid dumbbell model, can be given as
T
Zrot = -() ,
a

(18)

where a is the so-called symmetry factor to differentiate the heteronuclear molecules


(NO), a = 1, and homonuclear molecules (N2' O 2 ), a = 2 [14]. ()r is referred to as the
characteristic rotational temperature. I is the moment of inertial about its principal
axes.
From spectroscopic analysis, the characteristic rotational temperature of air mixture is found to be merely a few degrees above absolute zero (2.07K for O 2 , 2.86K for
N 2 ) [14]. A few intermolecular collisions are needed to achieve thermodynamic equilibrium. For most hypersonic applications, the internal translational and rotational
motion are considered to be fully excited, and will contribute 3/2RT and RT to the
internal energy, respectively. The term "fully excited" state really means that the
energy linearly increases with temperature.
Vibration and dissociation are closely related as a sequence of dynamic events.
The description of these two internal motions are much more complicated than the
internal excitations discussed previously. The simplest model of a vibrating diatomic
molecule is the harmonic oscillator. Wave mechanics requires that an oscillator has
a non-zero energy ground state, Ev = (i + 1/2) hv where i is a non-negative integer
and v is the fundamental vibrational freqency of the molecule. However, this model is
inconsistent with the fact that the diatomic molecule will dissociate. Individual atoms
behave as free particles beyond some value of the separation distance between them,
the diatomic molecule becomes dissociated. It behaves like an harmonic vibrator [14],
for which a crude model of the ionic bonding, homopolar bonding, and Van Der Waals
binding is usually adopted. The vibrational partition function and the energy of a
diatomic gas based on the simple harmonic oscillator are [11]
1
Zvib

1- exp ~

(19)

Computational Methods in Hypersonic Aerodynamics 91

Ev =

RBv
----:(,...--,)c--exp ~ - 1

(20)

where Bv = hv / k, is the characteristic temperature for vibration. For ideal dissociating


gas, Lighthill's model [26,27]' which is derived for a symmetric diatomic gas, has been
applied to simulate the non-equilibrium hypersonic flows past blunt bodies [20,28].
Numerical solutions, based on Lighthill's model, yielded reasonable results for detailed
flow structure and rate of heat transfer only for small departure from equilibrium
conditions.
The vibrational characteristic temperatures of air mixture are high; 2270K for
O2 , 3390K for N 2 , and 2740K for NO. The counterparts for the dissociation are
even higher; 59,500K and 113,000K for Oxygen and Nitrogen respectively [11]. In
order to achieve these energy states, a large number of collisions between particles
and radiative exchange are required which often leads to the non-equilibrium or the
relaxation phenomena.
Two speculations on the mechanism of dissociation are commonly discussed. The
climb-vibrational-ladder model, in which transfer from the bounded state to free atoms
is assumed to occur at the upper vibrational levels only. In contrast, the direct
transition model assumes a single violent collision to produce dissociation [14]. A
vibrational rate equation has been derived from the Landau- Teller collisional probabilities [11,14,16].

DEv
Dt

(21)

where E~ denotes the equilibrium value, and T is called the relaxation time, a function
of pressure and temperature. In theory, the Landau-Teller formulation is valid even
for large departure from equilibrium, but in practice the range of validity is limited.
A part of the limitation is imposed by the harmonic oscillator model which is only
accurate in the lower vibrational quantum levels. Again, the basic premise of this
formulation rests on the vibrational ladder climbing mechanism. Energy transition
is permitted only to the nearest energy level. This model has exhibited discrepancy
when compared with experimental facts [15,17]. Research efforts by Park et at. [18,29]
have attempted to bridge the prediction shortfalls by introducing a two temperature
model.
The contribution to thermodynamic properties from the excited atomic and electronic states is probably negligible for temperatures below 10,000K. This assertion is
supported by the fact that the next electronic energy level lies far above the ground
state [14]. The electronic partition function is naturally grouped into degenerate
levels [11].
Ze/

= 90

~
+ 79;

exp (

Be;)
-T

(22)

Again, Be,; denotes the corresponding characteristic temperature of electronic excitation. The only electron internal energy is associated with its spin, and contains
a degeneracy of two. The equilibrium composition of an ionized air mixture can be
analyzed by the the law of mass action, like that of dissociated air. The Saha equation
has been derived to yield the degree of ionization [11,14,30].

92 Computational Methods in Hypersonic Aerodynamics


It may be noted, that all the partition functions, Zrot, Zvib, and Zel are independent of volume V. This is a consequence of the assumption of weakly interacting
particles, for which the permissible internal quantum states of the particles have no
relation to the geometry of the system [11,14]. The only contribution to the pressure is from the translational partition function. According to Dalton's law of partial
pressure, the pressure of the system is the sum of all partial pressures of the species
present. Since the energy of different internal degrees of excitation is also additive,
the total internal energy of a system is found by summing over all the internal modes
and over all the species within the system.

Chemically Reacting Equations System


For air mixture at a temperature below about 9000K, the only species present in
sizable amounts are O 2 , N 2 , 0, N, and NO. All the forward chemical reactions,
including dissociation are endothermic, and not all subreactions are linearly independent. Our understanding of the chemical kinetics is based on the intermolecular
collision process, therefore the reaction induced by radiation will be excluded [11]. For
finite rate chemical reactions, the concept of local equilibrium must be used in the
computational procedure. The relaxation time scale in a non-equilibrium situation is
thus not unique. For this reason, the designation of a chemically reacting flowfield
as equilibrium, frozen, and relaxation by comparing the relative time scales between
molecular interaction and organized gas motion lacks rigor. It can be used only as a
convenient way to simplify analysis in applications.
The chemical reaction by a collision process contains reactant, activated complex,
and product. The transitory compound is formed by accumulating a sufficient amount
of activation energy for reaction to occur. This amount of energy, required to move
the reactants over the energy barrier in order for the reaction to begin, is usually
greater than the standard heat of formation. Conditions of the chemical reacting
system affecting the rate of chemical reactions in macroscopic states are numerous,
such as concentrations of the chemical compound, temperature, pressure, and the
presence of a catalyst or inhibitor. On a microscopic scale, the energy of relative
motion of colliding molecules, and relative orientation of the molecules at collision,
and the so-called steric factor are insoluble problems in quantum mechanics [11,14].
The chemical reaction can be expressed as
,,",'

,,","

(23)

J1i X i = L J1i X;

,-

where the J1~, and J1~ are the stoichiometric coefficients for reactants and products,
respectively. These coefficients are integers to denote the number of molecules that
partake in the reaction. Xi is the mole concentration of the reacting species.
The Law of mass action is actually an empirical formulation confirmed by numerous experimental observations. The law states that the rate of appearance or
disappearance of a chemical species is proportional to the products of the concentrations to a power equal to the corresponding stoichiometric coefficient, IIi-

,
Rj,i = /{j,i (T) II (Xit

(24)

Computational Methods in Hypersonic Aerodynamics 93


Similarly, rate of the reverse chemical reaction is given as
II

Rb,i

= Kb,i (T) II (Xi)";

(25)

where Kj,i and Kb,i are the so-called reaction rate constants which are functions of
temperature only and independent of the concentrations of the reacting species. The
net rate of change in concentration of any species Xi is the combined result of the
individual, elementary processes of single, simple-step reactions. In a complex overall
reaction consisting of a number of elementary reactions, the rate of reaction is often
determined by the slower rate of an elementary process. In this case, stoichiometric
coefficients may not necessarily govern the overall reaction rate.

dX' ) = ("
f ) Kj)'II(X;)V;,)
f
v-v
dt.
t,l
t,l
(I

+ (fv-v.
Kb'II(X)V;,)
't,l'}
II

II

t,}

(26)

where the subscript j, denotes the jth forward or backward reaction. In order to
apply the law of mass action, the rate of reactions is assumed to be independent
of whether or not the system is in equilibrium [11,14). This commonly accepted
assumption is consistent with the local equilibrium approach. The rate constant is
II

given as Kc,j = Kj,j/ Kb,j = II (X;)"i,j /II(X i )";,) which is considered to be a function
of temperature but not of the non-equilibrium variables. The relationship is equally
applicable to the dissociation-recombination reaction, as well as any simple, singlestep reaction (11). Only when the gas mixture is in chemical equilibrium, the forward
and reverse reactions are in dynamic balance, and the net rate of change in compostion
vanishes.

dXi) =
( dt.
)

(V~f.
_ v' .) K j
',)
I,)

.
,)

[II (X),,:,) __I_II (X)<)]


K.

(27)

CJ

In the macroscopic conservation formulation, the rate of chemical reaction is given in


terms of the mass concentration, rio
The so-called rate constant has been experimentally recognized as

K c,j = C exp ( -

I;~

(28)

This is the well-known Arrhenius equation, and the constant fa is often refered to as
the Arrhenius activation energy (11). The exponential term is also frequently denoted
as the Boltzmann factor. The Arrhenius Law simply states that only those molecules
that possess energy greater than a certain amount of activation energy will lead to
reaction. This equation has been modified to better fit to measurements as
KC,T = CTTI exp ( -

I;~)

(29)

Sometimes CTTI is called the collision frequency to reflect its root in statistical quantum mechanics. Unfortunately, the reaction rate constants are known with a large,
varying degree of uncertainty. At present, it is an intensive area of research. The
works of Park [29,31) are used as the bearer of standards for high temperature air.

94 Computational Methods in Hypersonic Aerodynamics


Energy Transfer by Radiation
The radiative mechanism of energy transfer is the direct consequence of shifting atomic
or molecular energy levels of a high temperature gas [11]. The transitions between
energy levels of a gaseous medium lead to the emission and absorption of photons.
Interaction between photons, motion of gas, and solid surfaces is reflected by the
scattering, emission and absorption of electromagnetic waves. Under equilibrium conditions, the energy absorbed and emitted by the gas must be balanced at every frequency. The temperature obtained from a Maxwell distribution will determine the
energy levels and populations of all components involved. In radiative equilibrium,
absorption and emission are accurately related by the Kirchhoff's law [11].
Under non-equilibrium conditions, the energy level population must be determined
first, then the absorption and emission coefficients can be calculated according to local
conditions. The equation of radiation transfer is easily recognizable as a special case
of the Boltzmann equation with the vanishing acceleration of photons [11].

a(nf) + Ci~ (nf) = [~(nf)]


at

aXi

at

E,A,S

(30)

where the photon number density and distribution function are denoted by nand
j, the velocity vector by Ci . E, A and S refer to the contributions by emission,
absorption and scattering respectively. To understand mechanisms of emission and
absorption of radiation in a gas requires knowledge of molecular and atomic spectroscopy. Atomic radiation arises from the bound-bound, bound-free, and free-free
transitions. Molecular radiation can trace their origins from the shifting of energy
levels between internal degrees of freedom. Since the radiant energy flux is directional and contains a range of frequencies, the energy transfer must be found in a
detailed description of the radiation field. In all, radiative intensity in one direction
is independent from any other direction, and is different from frequency to frequency.
These areas of special expertise are considered to be outside the scope of the present
discussion. However, several classic simplifications have been developed to evaluate
the radiation heat transfer and might be useful. In the Planck limit, emission is dominant, the optically thin gas approximation reduces the problem of just evaluating the
Planck absorption coefficient [11].
(31 )
where a denotes the Planck coefficient and a the Stefan-Boltzmann constant. Under
conditions of small departure from radiative equilibrium, the heat flux depends on the
first spatial derivative of the local temperature. The radiant heat flux can be obtained
by the Rosseland Diffusion approximation.

16aT3
qr = - - - VT
3ar

(32)

where a r denotes the Rosseland mean free path for radiation. Finally, the Grey-gas
approximation is also used to estimate the radiant energy flux. Since this approximation is not a rigorous average over the range of frequencies, accurate results cannot
be anticipated [11,18,32].

Computational Methods in Hypersonic Aerodynamics 95

Several approaches have been developed to calculate absorption and emission under non-equilibrium conditions. Two of the most frequently adopted methods are
the correction factor and the detailed line-by-line calculation over all frequency spectra [18,32,33). The correction factor method utilizes a simple equilibrium step model
for absorption, the effects of non-equilibrium population density are appended into
the calculation by a correction factor [33). On the other hand, the line-by-line calculation of spectra is capable of fully accounting for all absorption phenomena. Park et
al. [18) have found that under earth reentry conditions, the radiation emission from
non-equilibrium region is less intense than that from the equilibrium domain.
Macroscopic Conservation Equations
The complete conservation equations of the multi-component reacting system for hypersonic flows can now be written as:
The species conservation equations:

apai
at
+ \1. ( pai u + paiui )=.ri

(33)

where r denotes the rate of production. For an N number species system, only
N - 1 species conservation equations are required. Anyone of the N equations may
be replaced by the global equation of continuity for the mixture. In practice, all
species conservative equations are usually computed and verified by the balance of
the total number of atoms in the reacting system. The global continuity equation
can be obtained by summing over all species conservation equations and by noting
EPiUi = 0, and Eri = O.
(34)
which is the most fundamental principle of Newtonian mechanics; matter can be
neither created nor destroyed.
The conservation of momentum equations:

apu

-at + \1. (puu - f)

(35)

= p L..-fi

The conservation of momentum equation is the only vector equation in the system,
and it is the Newton's second law of motion. The nonlinear convective transport is
represented by the dyadic, the so-called inviscid term of the system. The shear stress
term (1') is a second rank tensor and can be given as

(36)
where A and Jl are the bulk and molecular viscosities respectively, and j is the identity
matrix.
The conservation of energy equation:

pe
aa
t

+ \1. (peu + q+ u.f) + Q =

p Ladi (u
.
I

+ Ui)

(37)

96 Computational Methods in Hypersonic Aerodynamics

It can be written in a directly usable form as:

aapet + \1. (peu + u.r -

k \1 T

+ p Laihiui + liT) + Lii~hr =


i

p Ladi. (u

+ Ui)

(38)
where e is the total energy and ~hi is the standard heat of formation of species i.
The system of equations will be formally closed when all the auxiliary equations for
transport properties, state variables and chemical reaction rate are prescribed. The
description of the differential system is completed only by incorporating all appropriate initial and boundary conditions.
In numerical simulations of hypersonic flows, this system of conservative equations
is usually written in strong conservation form.
(39)

EVOLUTION OF NUMERICAL PROCEDURES


Commonly Adopted Algorithms in CFD
The most widely used numerical algorithms in computational fluid dynamics probably can be attributed to MacCormack [34,35]' Beam-Warming [36], Jameson [37),
and more recently Thomas et af. [38,39). A significant number of variations and modifications of these procedures have been attempted to improve numerical efficiency,
accuracy, and range of applicability [23,35). However, some inherent limitations in
numerical analysis still need to be alleviated. Applying discrete methods to solve
highly nonlinear aerodynamic problems requires appending some form of numerical
dissipation, either implicitly or explicitly, into the complete algorithm to maintain
numerical stability. This added numerical dissipation must be sufficient to suppress
the high frequency wave cascade which leads to serious numerical instability, without
overwhelming the physics by generating an unacceptable error level. This requirement
in hypersonic applications, which involves extreme shock wave jump condition, is a
formidable challenge. An appropriate balance is usually not obtainable. In numerical
convergence process, the allowable temporal or spatial increments or both have to
be severely reduced to sustain the computations. Numerical efficiency degradation
is unavoidable. Even if convergence is reached, the shock wave is only discernible
by interpreting results from several grid points striding the discontinuity [28,40]' Figure 3. This deficiency cannot meet the numerical accuracy criterion, especially when
chemical physics non-equilibrium phenomena are of interest [41,42). The large variation of scales of fluid motion and chemical kinetics is reflected by the more complex
eigenvalue structure of the governing equations. This complete equation system of
hypersonic flow is very stiff for numerical simulation.
Recent progress in basic algorithm development aimed to overcome the aforementioned difficulties can be summarized into the areas of split flux and total variation
diminishing (TVD) schemes [39,43,44,45,46,47,48). All these methods have been developed for solving hyperbolic partial differential equations for which all eigenvalues
and characteristics of the system are real. Since the given initial values, together

Computational Methods in Hypersonic Aerodynamics 97


with any possible discontinuities are continued along characteristics, the solution of a
hyperbolic differential system is not necessarily an analytic function.
The rational approach of a shock-capturing method is mostly limited to the integral form of conservation equations via finite volume formulation. Although all the
TVD, flux vector splitting, and flux difference splitting schemes eventually are resulting in an approximation to the Reimann problem, the geneses of their development are
subtly different. The basic idea of flux vector splitting and flux difference splitting is
derived from the eigenvalue structure of the system of equations [39,43,44,46,48]. The
variations between flux vector splitting and flux difference splitting are differences in
numerical procedure based on the differentiability of the split fluxes [39,44,48]. The
TVD schemes, on the other hand, are developed according to the mathematical theory
of shock waves by Lax [49] for a scalar conservation law in which the total variation of
a solution can never increase, thus the discrete total variation shall not increase. However, in recent years, this faint line of difference in incorporating inherent information
from equations by these schemes seems to be disappearing. The most important issue
rests on the fact that the physical realism is effectively introduced by the Reimann
solver, exactly or approximately, into numerical analysis [44,48].
Split Flux Schemes
The basic idea of the flux splitting methods is to utilize the information of directional
signal propagation and invariant properties of characteristics of a hyperbolic partial
differential equations system. Therefore, these numerical schemes [39,43,44,46] are
directly tied to the eigenvalues and eigenvectors of the equation system. Since pertaining information is built in by the compatibility conditions, the implementation of
initial or boundary conditions is ensured to be both physically correct and mathematically well-posed. Furthermore, these methods can be classified in general as upwind
methods which have the advantage of being naturally dissipative, unlike in spatially
central difference schemes. No separate spatial dissipation terms are needed to maintain numerical stability. Unfortunately, the split flux methods rely on the integrating
procedure of a decoupled form of the compatibility relationships for multi-dimensional
equations, and require linearization of the equation system. In application, the procedures are nonunique, and are uncertain as to how precise these schemes need to
be [48]. Multi-dimensional problems are solved as multiple one-dimensional in each
direction, the possible bias of directional flux transport is neglected.
As an illustration for the genesis of these methods, the eigenvalue and the eigenvector of a two-dimension Euler equation will be derived to delineate the basic concept.

au aF aG _ 0
at + ax + ay where

u={

{~ }
pe

={

PuS
(pe

+ P) u

}G

={

(40)

pi:

(pe

(41)

+ P) v

It is well known, that the flux vector of the hyperbolic system of equations is a
homogeneous function of degree one in the elements of U [36]. By means of chain-rule

98

Computational Methods in Hypersonic Aerodynamics

differentiation, the system of equations may be expressed as

au

au

au

at + A 8x + B 8y

= 0

(42)

where, A = aFlau, and B = 8GI8U are the Jacobians of fluxes F and G.


The eigenvalue and eigenvector of the system of equations are found by manipulating the Jacobian with a similar matrix transformation. Since eigenvalues and
eigenvectors are independent of the transformation, if one can find a diagonal matrix of the equations, the coefficient matrices A and B can be expressed in diagonal
form in terms of eigenvectors in each spatial dimension. An inherent weakness of
all flux splitting methods originates from our inability to simultaneously diagonalize
the coefficient matrix for more than one-dimensional space. Therefore, the invariant
quantities are exchanged accurately in the direction normal to the control surface
between control volumes [48J. The fluxes parallel to the control surface are resolved
by uncertain approximations.
The eigenvalue of this system is obtained by solving the secular or the characteristic equation

u~'\ u~'\ ~
o
o

1~P ]=0

u-'\
0
0
u - ,\

,p

(43)

The eigenvalues are found to be ,\ = {u, u + c, u - c, u} where c is the speed of sound,


c2 = ,pip
The associated eigenvectors are obtained by Gauss elimination to reduce the
matrix,(A - U) = 0, into the Hermit normal form. In spite of the multiplicity of
one eigenvalue, the matrices A and B still have a very simple representation, the
so-called canonical representation or Jordan form. In the process of elementary row
elimination to reduce the matrices, an arbitrary basis has been assigned to construct
linearly independent eigenvectors. For reasons of simplicity, the values of unity are selected. The non-singular, similar transformation matrix of A is formed by containing
the eigenvectors as its columns

10]

1/C
1/c2
5 =
1I pc -1 I pc 0 0
0
0 1
[ o
1

(44)

0 0

If the left-hand inverse of the similar matrix 5 can be found, then the diagonalized
matrix A or B is realizable.

(45)
1

5- is obtained by a matrix inversion process

10
5 =

[ oo

pc/2
0
-pc/2

(46)

Computational Methods in Hypersonic Aerodynamics

99

The relationship between the diagonalized coefficient matrix and the characteristic
variables is derived in the following

aU

aU

at + AaX =
aU
at

A=

SAS-1 aU
aX

(47)

=0

(48)

~ u~c ~
[

o
o

0
0

u
0

Multiply equation (47) by the left-hand inverse of S, to get

S-l aU
at

AS- 1 aU =
aX
0

(49)

Assume S-l is a slowly varying local value function which allows the matrix to be
brought inside of the differential operator. This is the linearized approximation usually
needed in the differencing formulation to yield

Ow

ow

- + A - =0
at
ax
'

(50)

The system of one-dimensional characteristic equations are uncoupled:

~ (p at

P) + u~
(p - P) = 0
ax

c~

c~

(51 )

(52)
(53)

(54)
The subscript 0 indicates that the variables are evaluated at local conditions. Similarly, the system of characteristic equations for the other directions can be derived.
These equations simply show that along trajectories of dx = (u + c)dt, dx = udt, and
dx = (u - c)dt, the associated characteristic variables ws, are invariant. Equations
(50) through (53) are decoupled from each other, and are scalar equations. In this
form and with appropriate boundary conditions, it is the Reimann problem. The
purpose here is to demonstrate that all split flux schemes are based on characteristic
theory.
Since the decoupled system of equations in each spatial dimension is not necessarily easy to solve numerically and requires additional addressable memory for
characteristic variables in every spatial direction, a wide variation of flux vector splitting or the flux difference splitting methods were developed [39,43,44,46,48]. However,
a unique feature is shared among all flux vector splitting and flux difference splitting

100 Computational Methods in Hypersonic Aerodynamics


methods. At each time step, all solution procedures require the evaluation of a windward spatial difference according to the signs of eigenvalues. This discretization is
the direct consequence of the need to regulate flux exchange between adjacent control
volumes, and the orientation of information flow is controlled.
In flux vector splitting, the flux vector is split as

(55)
to separate the positive and the negative eigenvalues. Since the coefficient matrix
A is related to the flux vector by the homogeneous of order one relationship to the
dependent variables, F = AU. The coefficient matrix A, can be split to account
for the correct orientation of signal propagation. For the flux contribution from the
characteristic which has a positive slope (positive eigenvalue), backward discretization
is used. Accordingly, forward differencing is used for the flux with a negative slope or
negative eigenvalue.

Un+1 = un _ b.t (F+(W') _ F+(U n ))


,
'b.x
'
,-I

(56)
(57)

which is the sum of the total flux calculation F(U) = F+(U) + F-(U), and is the
numerical equivalent of requiring only the physically correct information exchange
between the adjacent control volumes.
The flux vector split algorithm may be extended to second-order by appropriate
extrapolation of information to the control surface. When the split fluxes are extrapolated, the method is called flux differencing. A better method is to use the MUSCL
approach [39]. In this procedure, the flow variables, either in conserved or primitive
form, are obtained on both sides of an interface by extrapolation. The extrapolation
process is then followed by the splitting of the fluxes.

8F..X

,-

F+(U-:-+ 1 )
'2

F+(U__ 1 )
'2

+ F-(U-:-+ 1 ) '2

b.x

F-(U __ 1 )
'2

(58)

where,

U- - U. ",Uj
j+t - ,+ '/-'.

Uj -

(59)
(60)

The second-order approximation of the flux splitting procedure requires some form of
a 'limiter' (<p) to prevent the occurence of under- and over-shoots of numerical results
immediately adjacent to discontinuous surface, and to reduce the solution accuracy
to first order.
Two of the more popular flux vector split approaches are originated by Steger et
al. [43] and Van Leer [46]. The former approach contains a non-differentiable flux
at sonic and stagnation points, where the eigenvalues either vanish or change sign.

Computational Methods in Hypersonic Aerodynamics

101

Numerical inaccuracies at these isolated points are manifested by the appearance


of glitches. Since both aforementioned methods turn out to be quite dissipative in
simulating viscous flow-fields, several modifications have been proposed to overcome
the noted deficiency [50).
Perhaps the most widely used MUSCL-based method is the flux difference split
method of Roe [44, 48). The flux formula for the Euler equations is obtained by
solving the linearized Riemann problem exactly, in such a manner that the jump
conditions are satisfied. The linearization process, however, introduces a complication:
the entropy inequality across the shock may not always be met. Therefore, artificial
dissipation must be appended to ensure physically meaningful solutions. Applications
in transonic and hypersonic flows have validated the superior quality of results using
the Roe's flux difference splitting scheme. One may conclude that the best numerical
procedure seems to be the one that is most closely associated with the local flow
properties. In a sense, it reflects its root in the characteristics of partial differential
equations, which is a point function of a field, according to differential geometry.
In hypersonic applications, the split flux techniques have also been applied to the
governing equations including the chemical kinetics effects by CoUela et at. [51), Candler [52), and Grossman et at. [53). In the work of Grossman et at. [53), both the flux
vector split and flux difference split formulations of the inviscid equations involving
chemical reaction, and energy relaxation processes are developed in a fully coupled
structure. Candler and MacCormack [52,62) in fact, pioneered the characteristic-based
method to solve the non-equilibrium hypersonic flows using Navier-Stokes equations.
Since the details of these efforts are rather lengthy to be included here, interested
readers may consult these references.
Total Variation Diminishing Scheme
The original notion of total variation diminishing (TVD) schemes is based on the TVD
property of solutions to homogeneous scalar hyperbolic conservation laws [45,47,49,54,
55).
0
(61 )

au al _
at + ax -

TV(u)

L: I~:I

00

(62)

dx;
-00

For this initial value problem, the total variation of consecutive solutions shall never
increase: TV(u n + 1 ) < TV(u n ). Therefore, solutions of TVD schemes have the unique
feature of not containing spurious oscillations when applied to one-dimensional nonlinear scalar hyperbolic conservation laws and constant coefficient hyperbolic systems [45,47). Since the applications of TVD schemes are mostly concentrated on the
shocks and contact discontinuities, weak solutions must be admitted. The piecewisesmooth weak solution of the conservation equation only satisfies the governing equation in two domains which are separated by the Rankine-Hugoniot jump. However,
weak solutions of the conservation law are not uniquely determined by the associated
initial data [47,55). The physically relevant solution is ensured by implementing a
numerical flux function as an entropy inequality criterion. The flux function is sometimes referred to as the numerical viscosity term, but in the modified-flux approach

102 Computational Methods in Hypersonic Aerodynamics


this function is used to achieve high resolution TVD schemes [45,47,54,55].
Extension of the scalar TVD scheme to Euler equations has some difficulty, since
unlike the scalar case, the total variation of the solution may increase at moments of
waves interaction [47,55]. In concept, this extension can be accomplished by defining a
local system of characteristic fields at each point. This procedure is not unique, since
it is the same as that of the flux vector split or the flux difference split previously
discussed.
The main mechanisms for maintaining the TVD property of a high-order scheme
are the flux limiter and the slope limiter. These limiting procedures are designed
to control either the magnitude of the flux function or the gradient of dependent
variables. High resolution TVD schemes all have automatic feedback mechanisms to
control the amount of numerical dissipation. Therefore, high-order accuracy will be
used to mean high order accurate almost everywhere but not at the discontinuity. The
symmetric or centered TVD schemes could become upwind biased or vice versa, by the
construction of a flux limiter. The clear definition of upwind and central differencing
approximation no longer exists for flux modified TVD schemes [47,55].
For the two-dimensional Euler equations in the Cartesian frame, equation (40),
the second-order symmetric TVD algorithm can be given as [47]

Ut+l
+ .6..6.tx (pnV.
_ F,:,+;1
.) + .6.t
,)
>+2"')
>-2"')
.6.y

(0':+11 >')+2"

G~+l1)
= ut).'
>')-2"

(63)

where the flux functions are

Fi+~,j
Fi_},j =

~ (Fi,j + Fi+1,j + Ai+~<I>i+~)


~ (Fi,j + Fi- 1,j + Ai_~ <I>i_~)

Gi,j+~
Gi,j_~

~ (Gi,j + Gi,j+1 + Bj+~<I>j+~)


~ (Gi,j + Gi,j-1 + Bj_~<I>j_~)

(64)

where A and B are the matrices of right eigenvectors of the Jacobian matrices 8Fj8U
and 8Gj8U, respectively. The matrices A and B are evaluated using Roe-averaged
variables. <I> is a vector that contains the first-order diffusion and second-order correction terms [55]. A typical element of <I> is
<I>i = -'Ij; (Ci - Qi)

(65)

Ci is the local characteristic speed of the considered characteristic variable, and 'Ij;
provides the entropy correction to enforce the selection of the physical meaningful
solution. In general, TVD schemes always converge, and their limit is the physical
weak solution [47,55].
All TVD schemes [45,47,54,55] employing a flux limiter satisfy the TVD property
for the hyperbolic scalar conservation laws. This feature ensures that the integration procedure is monotonicity preserving. However, these methods are not globally
second-order accurate [47,48]. The mechanism that eliminates spurious oscillations
in the solution also reduces the numerical accuracy locally to first-order at extrema

Computational Methods in Hypersonic Aerodynamics

103

and discontinuities. In any event, the TVD, flux vector split, and flux difference
split schemes have been very successfully applied to hypersonic problems. In solving the Euler equation, the improvement in shock wave definition over that of the
conventional methods is clearly demonstrated for both the single bow shock structure, Figure 4, and the complex shock-on-shock simulations [56]. Along the line of
symmetry, these higher-order resolution schemes uniformly capture the bow shock
waves between two adjacent cells which represent the best achievable by both split
flux and TVD schemes. For the more complex shock-on shock problems, these numerical results exhibit equally impressive improvements. However, quantification of
these results is difficult to obtain, and thus is not included in here.
CURRENT METHODOLOGY
Rarefied Gas Regime
The Burnett equations are the natural extension of the Navier-Stokes equations for a
rarefied gas from the development of Boltzmann solutions [12,24,57,58]. Near translational equilibrium, the Burnett equations are a higher order approximation beyond the
Chapman-Enskog expansion [12,58]. Another formulation for rarefied gasdynamics,
the thirteen-moment equations, was derived by Grad [59] through a truncated expansion for distribution function identifiable with that of Chapman and Enskog [57].
These two systems of equations for rarefied gas are formally equivalent to a Maxwell
gas [12,57], and in a state of small departure from equilibrium, the thirteen moment
equations can be reduced to the Burnett equations [57]. Cheng et al. [24,57] have
successfully developed a shock-layer theory based on the thirteen-moment equations
for rarefied hypersonic flows and achieved comparable results to that of the DSMC
method for hypersonic flow over a flat plate at high angle of attack.
The basic approach of the gas kinetic theory is based on the premise that if
the nature of collision process were known, the motion of gas could be described
by the Boltzmann equation with known initial conditions via a general concept in
statistics [11,12J. Numerical solutions to a statistical problem almost always rely on
the Monte Carlo methods. Here, as usual, any method that employs random numbers
may be referred to as a Monte Carlo method. The DSMC is gradually replacing the
Molecular Dynamics method and the Test Particle method to become the main stay
for rarefied gas simulation. In the DSMC method, the intermolecular collisions are
treated on a probabilistic rather than a deterministic basis and are required to assume
molecular chaos of dilute gas flows. The result is equivalent to a numerical solution
of the Boltzmann equations [60].
The DSMC method is a computationally intense particle simulation technique,
thus has very close ties with the advent of advanced computer architectures. In the
current state of development, validity and capability of DSMC methods have been
demonstrated by comparison with experiments and known numerical solutions [24,57,
60,61]. Future applications will be strongly influenced by the progress of computing
technology.
Continuum Flow Domain
The Navier-Stokes equations have a limited extension into the Knudsen number tran-

104 Computational Methods in Hypersonic Aerodynamics


sition region by means of the Maxwell's slip velocity and temperature jump conditions [9,12,25]. For a simple flow, this approximation yields a surprisingly realistic
result to that of the experimental data and Monte Carlo calculations [9,24]. However,
the applications to hypersonic flows are strictly restricted to a Knudsen number of
less than 0.1 [12,50]. Where the departure from the Maxwellian distribution becomes
marked, both the Chapmann-Enskog expansion and the Navier-Stokes equations cease
to be valid. However, the range of applications in hypersonic flows is still substantial
in critical stages of launching and reentering of any atmospheric or interplanetary
endeavors.
The most representative numerical algorithm in solving the Navier-Stokes equations for hypersonic applications is summarized by MacCormack and Candler [35,50,
52,62]. The basic approach by MacCormack [35,50] and among others [38,63,64] is
focused on the inclusion of a split flux formulation for the inviscid term of the NavierStokes equations and to solve the entire system of equations by the Gauss-Seidel
relaxation scheme. In principle, the split flux approach is based on the characteristic
theory, and is valid only for the hyperbolic equations system. The time-dependent
Navier-Stokes equations are known to be the incompletely parabolic type [23]. However, the numerical procedure has been very successfully applied to a wide range of
hypersonic flow problems including the chemical and thermal nonequilibrium phenomena.
In general, the implicit numerical procedure leads to the requirement of matrix
inversion. This numerical process requires a significant amount of calculations, and
the round-off error associated with a matrix of large dimensions can also be significant.
An alternative is to solve the system of equations by iterative or relaxation methods;
such as the Jacobi and the Gauss-Seidel algorithms. Each method involves transfer
of all the off-diagonal terms to the right-hand side of the equation to yield

OU; = b; -

Lif-j

a;joUj

a;i

(66)

where a;/s are the elements of the coefficient matrix, b;'s are the vector components
of the inhomogeneous terms. The difference between the Jacobi and Gauss-Seidel
method is that, unlike the former, the latter uses the most recently available value
of OU; to update solution without waiting for the complete set of new OU;. Applied
to elliptic differential equation, the Gauss-Seidel (Liebmann) scheme converges faster
than the Jacobi method (a special case of the Richardson scheme). The advantage
of the gain in numerical efficiency unfortunately also incurs an undesirable numerical
sweep bias [35,64]. Therefore, the Jacobi scheme is frequently adopted for the crossflow plane computation, and the Gauss-Seidel iteration is used in the streamwise
sweep to mimic the natural behavior of the flowfield development [64]. Convergence
of the solution is guaranteed if the coefficient matrix A can achieve the irreducible
form
A

[~~ ~3]

where Al is a square matrix, and the diagonal elements of matrix A must be dominant
over other entries.

Computational Methods in Hypersonic Aerodynamics 105


The relaxation procedure of MacCormack [35,50] for the three-dimensional N avierStokes equations, after discretization, can be outlined as following.
(67)
The inviscid terms are approximated by a split flux scheme. The current trend tends to
favor Roe's flux difference splitting procedure [44] in which the approximated Riemann
problem is solved exactly to satisfy the shock jump condition, and the flow properties
are obtained by extrapolation on both sides of the control surface [38,39,48]. The
viscous terms of the Navier-Stokes equations are approximated by spatially central
differencing. Therefore, all fluxes of the conservation equations are decomposed into
three distinct components.

(68)
(69)
(70)
Again the inviscid terms are linearized:

F~ + (~~): (U n+ un)

(71)

F~ +A~bU

(72)

F~ + (~~): (un+! - un)

(73)

F~

+ A~bU

(74)

Identical procedures are also applicable to the inviscid flux terms in two other spatial
orientations. The viscous terms may be similarly linearized:

Fn+

Fn+ bFn
v

Substituting the above developments into equation (67) to yield

(75)
Separating the inviscid and viscous terms, the above equations are expressible as

{I + ~t
~t

[! (A

+ A-) +

:y (B+ + B-) + :z (c+ + c-)]}bU +

ax + (ac)n
ay + (aH)n]
az
(76)
[axa (bF:;) + aya (bG~) + aza (bH:) ] = -~t [(aF)n

Note that the Right hand side represents the residual and may be evaluated by any
desired method. The treatment of the viscous terms requires some additional attention. First, the energy fluxes contributed by viscous effects are functions both of the

106 Computational Methods in Hypersonic Aerodynamics


dependent variables and its first-order derivatives. In this formulation, all the viscous
terms, that originate from the dissipation function, u. T, are permitted to lag in temporal evolution. Second, the shear stress tensor defined by the primitive variables,
V = {p, u, v, w, ed, must now be related to the conservative variables.

(77)
where, the coefficient matrix defines the relationship between the viscous flux and the
primitive variables [35,50,64]. The grouping of the conservative dependent variables
from the primitive variables is related by the following transformation Jacobian

.{ )~
p
pu

RV

pw
pe

1
u
v
w

0
P

0
0

pu

pv

p
pw

0
0
0
0

p
u
v
w

ei

)~ N-'l )
p
u
v
w

(7S)

ei

The typical flux component contributed by the viscous terms now can be given as

8Fv =

-Mxfj0 (N8U)

(79)

Xi

Substituting all the previous derivations into equation (76), we have

{I+!:It [!..-ox (A+ + A- - Mx~N) +!..-oy (B+ + B- - M~N) +


:z (c+ +c- - Mz O~i N)]} 8Un -!:It [( ~:) n + (~~) n + (~~) n]
OXi

OXi

(SO)

It may be interesting to note that although the left-hand-side of the implicit differencing of the viscous terms can be the thin-layer approximation to the Navier-Stokes
equations, all the viscous terms of the Navier-Stokes equations, as the driving force of
the implicit numerical formulation [35,50], must remain intact on the Right-hand-side
of equation (SO). The converged numerical solution to a steady asymptote will then
satisfy the exact macroscopic equation of motion.
According to MacCormack's algorithm [35,50] the time advancement of a numerical solution is given by the implicitly predicted change of the conservative variables
8U. This implicit increment is obtained by satisfying the explicit change of the NavierStokes equations,

!:lU.
{I+!:It [!..-ox (A+ + A- - Mx~N) +!..-oy (B+ + B- - M~N) +
!..ax (c+ + c- - M~N)]} 8U = -!:lU .
OXi

Ui~t1 = Ui~j,k

8Xi

+ 8Ui ,j,k

't,),

OXi

t,),

(S1)
(S2)

This linear algebraic system is factored in delta form and solved by combined GaussSeidel and Jacobi iterative sweeps [64]. In each spatial direction, the factored one

- - - - - - - _ .__.

Computational Methods in Hypersonic Aerodynamics

107

dimensional system equations can be rearranged to present the dependent variables


in the ascending order of index j as
(83)

Since the alternative backward and forward numerical sweeps are standard procedures
for a relaxation calculation, the detailed initiation and initial/boundary condition
implementations will not be repeated here [35,50,64].
In essence, MacCormack's explicit/implicit finite volume numerical method [35],
including a flux splitting option, has been demonstrated to be very effective for solving three-dimensional, Navier-Stokes equations. At the present time, it is one of the
most popular algorithms for solving hypersonic flow problems. The spatially secondorder accurate results, and the alleviated numerical directional bias by a sweep lagging
process have been achieved with a minimum degradation of numerical efficiency. As
a matter of fact, nearly an order of magnitude of reduction of computational resource was realized in comparison with earlier numerical procedures [40,64]. More
importantly, comparable numerical results have been achieved by this method with
much lower numerical resolution requirement [42,64]. The same numerical procedure
is equally effective to simulate the high temperature hypersonic flow, in which the
added complications of chemical and thermal non-equilibrium phenomena have introduced additional equations of chemical physics into the system [32,42,52,62].
To summarize, for hypersonic flows in the continuum regime, numerical simulations require a solving procedure for equations of motion and chemical physics. From
statistical quantum mechanics, the atomic and molecular structure and its macroscopic behavior in high temperature in equilibrium state are well understood [11,14].
Transport properties of gas media and radiative heat transfer in limiting conditions
are also known [11,13,25]. However, in hypersonic applications, the chemical and thermodynamic equilibrium state is rarely attainable. The basic physical phenomena are
evolving around the energy transport by both elastic and inelastic collision processes,
and by the shifting of elementary particles to different energy levels through radiation.
All these mechanisms require a wide range of time and length scales to proceed to
an asymptote, if one exists. At present, the non-equilibrium problem is studied on
the framework of the concept of local equilibrium. Little assurance can be assessed
as to what extent the concept of a small departure from equilibrium is still valid for
numerical simulation. Therefore, the pacing items in hypersonic research are hinged
on the progress in chemical kinetics and the statistical quantum mechanics for the
nonequilibrium phenomena.
In the continuum regime, the governing equations of three-dimensional hypersonic flows include five scalar components of the compressible Navier-Stokes for the
conservation of the mass, momentum, and the energy. Excluding the heterogeneous
ablating chemical reactions and the trace of noble gases, the dissociating and ionizing
Nt,
air may have eleven individual reacting species (N2 , O 2 , N, 0, NO, 0+, N+,
NO+, e+). In a state of thermal nonequilibrium, additional equations to describe energy exchange between internal degrees of freedom and a system of integro-differential
equations for radiation are also required. In the rarefied gas domain, the N avier-Stokes
equations may be applied with modifications of the surface boundary conditions to
reflect the transition from dominance of molecular collisions to interaction with solid

ot,

108 Computational Methods in Hypersonic Aerodynamics


surfaces [12,25]. The most recent approach, however, has moved into the direction
of higher order approximations to the Boltzmann equation [24,57], or by the direct
simulation Monte Carlo (DSMC) method [60,61].
At the present time, numerous and successful simulations of hypersonic nonequilibrium phenomena have been achieved by using a variety of numerical schemes
[28,41,42,52,62]. But we still have an arduous journey ahead of us in attaining the
technological maturation needed to understand and to predict the interdisciplinary
phenomena in hypersonic flows.
ACKNOWLEDGEMENT
The author appreciates deeply the support from Mr. E. Josyula, Drs. M.R. Visbal, T.S. Tavares and R.W. Newsome as well as the members of the Computational
Aerodynamics Group. The generous help and stimulating discussions with Dr. D.
Gaitonde are gratefully acknowledged.

References
[1] Tsien, H.S., Superaerodynamics, Mechanics of Rarefied Gases, J. Aero. Science,
Vol 13, 1946, pp 653-664.
[2] Hays, W.D. and Probstein, R.F., HYPERSONIC FLOW THEORY, First ed.
Academic Press, New York, 1959.
[3] Van Dyke, M.D., Application of Hypersonic Small-disturbance Theory, J. Aero.
Science, Vol. 21, 1954, pp 179-186.
[4] Hays, W.D., On Hypersonic Similitude, Quart. Applied Math. Vol 5, 1947, pp
105-106.
[5] Il'yusin, A.A., The Law of Plane Sections in the Aerodynamics of High Supersonic
Speeds, Prikl, Mat. Mekh. Vol. 20, 1956, pp 733-755.
[6] Taylor, G.I., The Formation of a Blast Wave by a very Intense Explosion, Proc.
Roy. Soc. London. Ser. A., Vol. 201, 1950, pp 159-186.
[7J Sedov, L.I., On Unsteady Motions of a Compressible Fluid, Dokl. Akad. Nauk
SSSR, Vol 47, 1945, pp 91-93.
[8] Stewartson, K. and Williams, P.G., Self-induced Separation, Proc. Roy. Soc.
Series A, Vol 312, 1969, pp 181-206.
[9] Gokcen, T. MacCormack, R.W. and Chapman, D.R., Computational Fluid Dynamics Near the Continuum Limit", AIAA 87-1115-CP, 1987.

[10] Cheng, H.K., Hypersonic Shock-Layer theory of the Stagnation Region at Low
Reynolds Number, Proc. Heat Transfer and Fluid Mech. Inst. Stanford Univ.
Press, 1961, pp 161-175.

Computational Methods in Hypersonic Aerodynamics 109


[11] Vinceti, W.G. and Kruger C.H., INTRODUCTION TO PHYSICAL GAS DYNAMICS, John Wiley and Sons, New York, 1965.
[12] Chapman, S. and Cowling T.G., THE MATHEMATICAL THEORY OF NONUNIFORM GASES, Cambridge Univ. Press, 2nd edition, 1964.
[13] Bird, R.B., Stewart, W.E. and Lightfoot, E.N., TRANSPORT PHENOMENA,
John Wiley and Sons, Inc., New York, 1960.
[14] Clarke, J.F. and McChesney, M., THE DYNAMICS OF REAL GAS, Butterworth & Co., London, 1964.
[15] Blackman, V., Vibrational Relaxation in Oxygen and Nitrogen, J. Fluid Mech.,
Vol 1, Part 1, 1956, pp 61-85.
[16] Landau, L. and Teller, E., Zurtheorie der Shallispersion, Physik Z. Sowjetunion,
B, 10, h.1, 1936, p 34.
[17] Herzfeld, K.F., The Rate and Mechanism of the Thermal Dissociation of Oxygen,
7th International Symposium on Combustion, Butterworth & Co., London, 1959.
[18] Park, C., Howe, J.T., Jaffe, R.L. and Candler, G.V., Chemical-Kinetic Problems
of Future NASA Missions, AIAA 91-0464, 1991.
[19] Taylor, H.S., Combustion Processes, Vol II, The High Speed Aerodynamics and
Jet Propulsion Series, Edit. Lewis, B., Princeton University Press, 1956.
[20] Josyula, E. and Shang J.S., Numerical Simulation of Non-equilibrium Hypersonic
Flows with Wall Catalytic Effects, AIAA 89-0462, 1989.
[21] Kendall, J.M., Wind Tunnel Experiments Relating to Supersonic and Hypersonic
Boundary-Layer transition", AIAA J. Vol. 13, No.3, March 1975, pp 290-299.
[22] Stetson K.F., Comments on Hypersonic Boundary-Layer Transition, WRDC-TR90-3057, Wright-Patterson Air Force Base, Ohio, September 1990.
[23] Shang, J.S., An assessment of Numerical Solutions of the Compressible NavierStokes Equations, J. Aircraft Vol 22, No 5, May 1985, pp 353-370.
[24] Cheng, H.K., Wong, E.Y., and Dogra, V.K., A Shock-Layer Theory Based on
Thirteen-Moment Equations and DSMC Calculations of Rarefied Hypersonic
Flows, AIAA 91-0783, 1991.
[25] Hirschfelder, J.O., Curtiss, C.F. and Bird, B.B., MOLECULAR THEORY OF
GASES AND LIQUIDS, 2nd printing, John Wiley & Son, New York, 1954, pp
514-667.
[26] Lighthill, M.J., Dynamics of a Dissociating Gas - Part I, Equilibrium Flow, J.
Fluid Dynamics, Vol. 2, 1957, pp 1-32.
[27] Freeman, N.C., Non-equilibrium Flow of an Ideal Dissociating Gas, J. Fluid Mech.
Vol. 4, 1958, pp 407-425.

110 Computational Methods in Hypersonic Aerodynamics


[28] Shang, J.S. and Josyula, E., Numerical Simulations of Non- equilibrium Hypersonic Flow Past Blunt Bodies, AIAA 88-0512, 1988.
[29] Park, C., Assessment of Two-temperature Kinetic Model for Ionizing Air, AIAA
J. of Thermophysics and Heat Transfer, Vlo. 3 No.3, July 1989, pp 233-244.
[30] Saha, M.N., Ionization in the Solar Chromosphere, Phil. Mag. Vol. 40, No. 238,
1920, p. 472
[31] Park, C., A Review of Reaction Rates in High Temperature Air, AlA A 89-1740,
1989.
[32] Candler, G. and Park, C., The Computation of Radiation from Nonequilibrium
Hypersonic Flows, AIAA 88-2678, 1988.
[33] Carlson, L.A., Approximations for Hypervelocity Nonequilibrium Radiating, Reacting, and Conducting Stagnation regions, AIAA J. of Thermophysics and Heat
Transfer, Vol.3, No.4, Oct. 1989, pp 380-388.
[34] MacCormack, R.W., The Effect of Viscosity in Hypervelocity Impact Cratering,
AIAA 69-0354, 1969.
[35] MacCormack, R.W., Current Status of Numerical Solutions of the Navier-Stokes
Equations, AIAA 85-0480, 1985.
[36] Beam, R.M. and Warming, R.F., An Implicit finite Difference Algorithm for
Hyperbolic Systems in Conservation-Law-Form, J. Compo Phys. Vol 22, 1976, pp
87-110.
[37] Jameson, A., Schmidt, W., and Turkel, E., Numerical Solutions of the Euler Equations by Finite Volume Methods Using Runge-Kutta Time-stepping
Schemes, AlA A 81-1259, 1981.
[38] Thomas, J.L. and Walters, R.W., Upwind Relaxation Algorithms for the NavierStokes Equations, AIAA 85-1501CP, 1985.
[39] Anderson, W.K., Thomas, J.L., and Van Leer, B., Comparison of Finite Volume
Flux Vector Splittings for the Euler Equations, AIAA J. Vol. 22, No.9, Sept.,
1986, PP 1453-1446.
[40] Shang, J.S. and Scherr, S.J., Navier-Stokes Solution for a Complete Reentry
Configuration, J. of Aircraft, Vol 23,No 12, 1988, pp 881-888.
[41] Josyula, E. and Shang, J.S., Numerical Study of Hypersonic Dissociated Flow
Past Blunt Body, AIAA J. Vol 29, No 5, 1991, pp 704-71l.
[42] Josyula, E., Gaitonde, D. and Shang, J.S., Numerical Simulation of Nonequilibrium Hypersonic Flow Using Roe Flux-difference Splitting Scheme, AIAA 911700, 1991.
[43] Steger, J.L. and Warming, R.F., Flux Vector Splitting of the Inviscid Gasdynamics Equations with application to finite difference Methods, J. Compo PhYsics,
Vol 40, 1981, pp 263-293.

Computational Methods in Hypersonic Aerodynamics 111

[44] Roe, P.L., Approximate Riemann Solvers, Parameter Vectors, and Difference
Schemes, J. Compo Physics, Vol 43, 1981, pp 357-372.
[45] Harten, A., High Resolution Schemes for Hyperbolic Conservation Laws, J.
Compo Physics, Vol. 49, 1983, pp 357-393.
[46] Van Leer, B., Flux-vector Splitting for the Euler Equations, Lee. Notes Phys. Vol
170, 1982, pp 507-512.
[47] Yee, H.C., Construction of Implicit and Explicit Symmetric TVD Schemes and
Their Applications, J. Compo Physics, Vol 68, 1987, pp 151-179.
[48] Roe, P.L., Characteristic-based Schemes for the Euler Equations, Ann. Rev. Fluid
Mech. Vol 18, 1986, pp 337-365.
[49] Lax, P., Hyperbolic Systems of Conservative Laws and the Mathematical Theory
of Shock Wave, SIAM Regional Series on Applied Math., Volll, 1973.
[50] MacCormack, RW. and Candler, G.V., The Solution of the Navier-Stokes Equations Using Gauss-Seidel Line Relaxation, Computers & Fluids, Vol. 17, No.1,
1989, p135-150.
[51] Collela, P. and Glaz, H.M., Efficient Solution Algorithm for the Reimann Problem
for Real Gases, J. Compo Physics, Vol 59, 1985, pp 264-289.
[52] Candler, G.V., The Computation of Weakly Ionized Hypersonic Flows in Thermochemical Nonequilibrium, Ph.D. Thesis, Stanford University, Calif. 1988.
[53] Grossman, B. and Cinnella, P., Flux-split Algorithms for Flows with Nonequilibrium Chemistry and Vibrational Relaxation, J. Compo Physics, Vol 88,
1990, pp 131-168.
[54] Harten, A., On a Class of High Resolution Total-Variation- Stable Finite Difference Schemes, Siam J. Num. Analy., Vol 21, 1984, pp 1-23.
[55] Yee, H.C. Warming, R.F., and Harten, A., Implicit Total Variation Diminishing
(TVD) Schemes for Steady-State Calculations, J. Compo Physics Vol 57, 1985,
pp 327-360.
[56] Kroll, N., Gaitonde, D., and Aftosmis, M., A Systematic Comparative Study of
Several High Resolution Schemes for Complex Problems in High Speed Flows,
AIAA 91-0636, 1991.
[57] Cheng, H.K., Lee, C.J., Wong, KY, and Yang, H.T., Hypersonic Slip Flows and
Issues on Extending Continuum Model Beyond the Navier-Stokes Level, AIAA
89-1663, 1989.
[58] Burnett, D., The Distribution of Molecular Velocities and the Mean Motion in a
Non-Uniform Gas, Proc. Lond. Math. Soc., Ser 2, Vol 40, 1936, pp 382-397.
[59] Grad, H., On the Kinetic Theory of Rarefied Gas, Comm. of Pure Appl. Math.,
Vol 2, No.4, 1949, pp 331-407.

112

Computational Methods in Hypersonic Aerodynamics

[60] Bird, G.A., Monte Carlo Simulation in an Engineering Context, Prog. In Astro.
and Aero.: Rarefied Gas Dynamics, Vol. 74, Part 1, Edited S. Fisher, 1981, pp
239-255.
[61] Moss, J.N., Bird, G.A., and Dogra, V.K., Nonequilibrium Thermal Radiation for
an Aeroassist Flight Experimental Vehicle, AIAA 88-0081, 1988.
[62] Candler, G.V. and MacCormack, R.W., The computation of Hypersonic Ionized
flows in Chemical and Thermal Nonequilibrium, AIAA 88-0511, 1988.
[63] Newsome, R.W., Walters, R.W, and Thomas, J.L., An Efficient Iterative Strategy
for Upwind/Relaxation Solutions to the Thin- Layer Navier-Stokes Equations,
AIAA 87-113, 1987.
[64] McMaster, D.L., Shang, J.S. and Gaitonde, D., A Vectorized Gauss-Seidel Line
Relaxation Scheme for Solving 3D Navier-Stokes Equations, AIAA 89-1958CP,
1989.

Computational Methods in Hypersonic Aerodynamics 113

MACROSCOPIC DESCRIPTION
(CONTINUUM MECHANICS)

MICROSCOPIC DESCRIPTION
(FREE-IIOLECULE DYNAMICS)
CHAPIIAN-ENSKOG EXPANSION
BURNETT EQUATIONS
THIRTEEN-MOMENT EQUATIONS

INTERACTIVE BOUNDARY-LAYER THEORY


THIN-LAYER N_S. EQUATIONS

STEADY
RELATIVELY SIMPLE
CONAOURATION

LIMITED DOMAIN
01
CONSEQUENCE

INERTIA FORCE
DOMINANT

SMALL DISTURBANCE

Figure 1: Hierarchy of CFD predictive methodology.

Moo =12
AlT = 30,480 m

-.... . . . . . . . DISSOCI.AT.I
.
...
N-G-G-A-S--------

Figure 2: Mach number contours about blunt body [28].

116 Computational Methods in Hypersonic Aerodynamics

Cv,tr

Cv,V
D

D.
eV
eV,s

l
g
h
h
hs
hv,s
H
I
kb,r
kj,r

lx, ly, lz

L/D

m x , my,

mz

ML
ML,INV
ML,vIS
ML,SRC

Ms
n

n
n x , ny, n z

n
Nr
Ns
p
q
Qrad

rjINv
rjvIs

drag coefficient
heat transfer coefficient
reference heat transfer coefficient
lift coefficient
specific heat for translational-rotational energy
specific heat for vibrational-electronic energy
base diameter
effective diffusion coefficient for species s
energy per unit mass of species s
mixture vibrational-electronic energy per unit mass
vibrational-electronic energy per unit mass of species s
total energ~ per unit mass of mixture
flux vector in Cartesian space
inviscid component of flux vector relative to cell face
viscous component of flux vector relative to cell face
enthalpy or altitude
enthalpy per unit mass of species s
vibrational-electronic enthalpy per unit mass of species s
total enthalpy per unit mass of mixture
identity matrix
backward reaction rate coefficient for reaction r
forward reaction rate coefficient for reaction r
unit vector tangent to computational cell wall
components of
in X,y,z directions, respectively
lift to drag ratio
unit vector tangent to computational cell wall
components of m in x,y,z directions, respectively
point-implicit Jacobian of flux terms
point-implicit Jacobian of inviscid terms
point-implicit Jacobian of viscous terms
point-implicit Jacobian of source terms
molecular weight of species s
iteration level index
unit vector normal to computational cell wall
components of n in x,y,z directions, respectively
number density
number of reactions
number of species
pressure
vector of conserved variables
radiative energy transfer rate
right-hand-side residual vector
relaxation factor used with inviscid Jacobian matrices
relaxation factor used with viscous Jacobian matrices

Computational Methods in Hypersonic Aerodynamics 117

Tv
u,v,w
U, V, W

X,Y,Z
Ys
a

/3slr
ir
10, to

eN
'TJ
7]v

Oe
OR
A
~
A
/l
/lsj
~,'TJ, (

p
p.
(j

1>
X
w

universal gas constant


backward reaction rate for reaction r
forward reaction rate for reaction r
nose radius
shoulder radius
matrix of left eigenvectors of A
arc length
arc lengths in ii,~ m directions
approximation to gradient of characteristic variables
function of s
time
translational-rotational temperature
vibrational-electron-electronic excitation temperature
velocity components in x,y,z directions
velocity component in the ii, ~ and m directions
Cartesian coordinates
mole fraction of species s
angle of attack relative to base plane normal
stoichiometric coefficient for reactants in reaction r
8pj8pE
stoichiometric coefficient for products in reaction r
8pj8Pr
parameters for defining minimum eigenvalue
eccentricity of nose
frozen thermal conductivity for translational-rotational energy
frozen thermal conductivity for vibrational-electronic energy
o for first-order, 1 for second-order approximation to inviscid flux
cone half angle in plane of symmetry
rake angle of base plane
eigenvalue of A
restricted eigenvalue of A
diagonal matrix of eigenvalues of A
mixture viscosity
reduced mass of species sand j
computational coordinates
mixture density
density of species s
cell face area
cross section for translational-vibrational energy exchange for species s
shear stress or relaxation time
8pj8p ev
dummy variable for ~, 7], or (
vector of source terms
mass production rate of species s per unit volume

118 Computational Methods in Hypersonic Aerodynamics

Wv

vibrational-electronic energy source term


cell volume

Subscripts
e
i,j, k

I,J,K
1
L
r

s
V
00

electron
indices of cell walls in the ~, 'f/, ( directions
indices of cell centers in the ~,'f/, ( directions
dummy index for a cell wall
dummy index for a cell center
reaction r or species r
speCIes s
vibrational-electronic
free stream

INTRODUCTION
The National Aeronautics and Space Administration's interest in viscous, hypersonic
flow field simulation has grown in recent years in anticipation of the design needs
for space transportation and exploration over the next three decades, e.g. Walberg
[1]. Aeroassisted space transfer vehicles will use the upper layers of planetary atmospheres in hypersonic aerobraking maneuvers. The National Aero-Space Plane,
with its supersonic combustion ramjet engines, is being designed to achieve hypersonic-speeds through the Earth's atmosphere in its climb to orbit. Various concepts
for a second-generation Space Shuttle are now being considered, as are concepts for
a space station crew emergency return vehicle. The external flowfield surrounding
such vehicles, as well as the internal flowfield through the scramjet engine, may be
significantly influenced by thermochemical nonequilibrium processes in the flow. Accurate simulations of these phenomena would provide designers valuable information
concerning the aerodynamic and aerothermodynamic character of these vehicles.
There are two major challenges to the simulation of flowfields in thermochemical
nonequilibrium around vehicles travelling at hypersonic velocities through the atmosphere. First, these simulations require modelling of the nonequilibrium processes in
the flow, frequently at energies where the models currently lack sufficient experimental or analytic validation. Nonequilibrium processes occur in a flow when the time
required for a process to accommodate itself to local conditions within some region
is of the same order as the transit time across the region. The equations and models
used in this chapter for non-equilibrium flow have been documented in Reference [2]
and were substantially derived from the work of Park [3, 4] and Lee [5]. The conservation equations are formulated within the context of the LAURA algorithm (Langley
Aerothermodynamic Upwind Relaxation Algorithm) [7]. Calibration and validation
of the physical models intrinsic to this code, which must deal with nonequilibrium

Computational Methods in Hypersonic Aerodynamics 119

chemical and thermal processes, curve fits of thermodynamic and transport properties
at high temperatures, and radiative energy transfer, was first discussed in Reference
[8]. Other code development and calibration programs (e.g. Candler [9, 10], Park
and Yoon [11], Netterfield [12] Coquel et al. [13]) are now in progress within the area
of viscous, hypersonic, reacting gas flowfield simulations. Still, these calibration and
validation studies do not fully validate the flowfield simulations at velocities characteristic of vehicles returning to earth from the Moon or Mars. These simulations
should be used to help design ground-based and flight experiments for obtaining the
necessary data by performing parametric studies which identify critical aspects of the
physical models and locations in the flowfield which are most sensitive to unknowns
in these physical models.
Second, because of the large number of unknowns associated with chemical
species and energy modes and because of disparate time scales within the flowfield,
these simulations require algorithmic innovations to maintain numerical stability and
fully exploit supercomputer resources. Numerical stability is maintained through an
implicit treatment of the governing equations. A great variety of implicit treatments
is possible. For problems in which only the steady state solution is required, one is
free to evaluate any element of the difference stencil at any iteration (pseudo time)
level which facilitates the relaxation process. In the most rigorous implicit treatment,
all variables in all cells are simultaneously solved at an advanced iteration level, requiring the solution of a linearized equation set involving nI J J{ equations where n is
the number of unknowns at a point and I, J, and J{ are the number of computational
points in the three respective coordinate directions. The various forms of factored implicit schemes and line relaxation methods sequentially solve equation sets involving
nI, nJ, and/or nJ{ variables. The point-implicit schemes, as utilized in the present
work, sequentially solve equation sets involving n simultaneous, linearized equations.
Further simplification is possible in chemical kinetic problems by linearizing contributions to the residual from only the source terms to alleviate problems of disparate
chemical time scales, resulting in methods which involve no matrix operations.
The essence of the point-implicit strategy is to treat the variables at the cell center
of interest implicitly at the advanced iteration level and to use the latest available
data from neighbor cells in defining the "left-hand-side" numerics. The success of this
approach is made possible by the robust stability characteristics of the underlying
upwind difference scheme. The basic algorithm requires only a single level of storage,
and numerical experiments show excellent stability characteristics, even when working
directly with the steady state equations (i.e., Newton iteration). Furthermore, the
algorithm is particularly amenable to generating solutions on unstructured grids [14]
and is efficiently implemented on vector or parallel processors [15]. Details of the
relaxation algorithm for the case of a gas in thermal and chemical nonequilibrium
are presented herein. Sample applications related to hypersonic flows over aero brakes
follow.

120 Computational Methods in Hypersonic Aerodynamics

FINITE-VOLUME FUNDAMENTALS
The integral form of the conservation laws applied to a single cell in the computational
domain is written

JJJ~7

dD

JJf .

JJJw

ii da =

dD

(1)

In Equation 1 the first term describes the time rate of change of conserved quantity
q in the control volume; the second term describes convective and dissipative flux f
through the cell walls; and the third term accounts for sources or sinks of conserved
quantities within the control volume. The third term is identically zero for perfect-gas
flows but is required for flows in chemical or thermal nonequilibrium.
The finite-volume approximation to Equation 1 for a general, unstructured grid
is written
I:

[ ~]
Of L +

IIh

""'
L.J f-m

- m
n",a

= [.W H0] L

(2)

m=l

where oq = qn+1 - qn, ot = t n+ 1 - tn, Ah is the number of faces of cell L


having volume D, and subscript m refers to cell face m with surface area am. The
quantity
is a unit vector normal to cell face m in a direction facing away from
the cell center. The dependent variable q is defined at cell centers. The independent
variables x, y, and z are defined at cell corners.
The finite-volume approximation to Equation 1 for a rectangularly ordered, structured grid is written

nm

oq D

[8t h,J,l(

+ [ ;+1

ni+1 a i+1

~+1

nj+I a j+1

+ [ fk+I

nk+1 a k+1

i-J
fk

niaj ]J,K

(3)

na']IY
J J
, \
i"hak

]l,J = [w D]l,J,K

A shorthand notation for Equation 3 that will be used throughout this paper follows:

L [;+1 . 11/+1 a/+1

; .

11/a/]

= [w D]L

(4)

/=i,j,k

Note in Equations 2-4 that uppercase integer variables I, J, J{, and L denote computational coordinates at cell centers, and lowercase integer variables i, j, I.:, t, and m
denote cell faces or cell corners. For example, ai,J,h" refers to the cell wall corresponding to indices 1-1/2, J, J{. (See Figure 1.) In the shorthand notation of Equation 4,
the integer variable I is used as a generic index for i, j, or k. This notation is convenient because most of the formulations for quantities at cell faces are independent
of the coordinate direction. The geometric quantities D, a, and 11 are easily derived
given the Cartesian coordinates of the cell corners. Details may be found in Appendix
A of Reference [6J.
The formulations that follow are based on a rectangularly ordered, structured
grid. A first-order-accurate formulation of the inviscid equations on a structured grid

Computational Methods in Hypersonic Aerodynamics 113

MICROSCOPIC DESCRIPTION
(FREE-IIOLECULE DYNAMICS)

.....CROSCOPIC DESCRIPTION
(CONTINUUM MECHANICS)
CHAPIIAN-ENSKOG EXPANSION
BURNETT EQUATIONS
THIRTEEN-IIOIIENT EQUATIONS
INTERACTIVE BOUNDARY-LAYER THEORY
THIN-LAYER N.S. EQUATIONS

STEADY
RELA TlVEL Y SIMPLE
CONAOURATION

LIMITED DOMAIN
01
CONSEQUENCE

INERTIA FORCE
DOMINANT

SMALL DISTURBANCE

Figure 1: Hierarchy of CFD predictive methodology.

= 12
AlT = 30,480 m

Mm

.........

_..

DISSOCIATING GAS

Figure 2: Mach number contours about blunt body [28].

114 Computational Methods in Hypersonic Aerodynamics

o.g
D

0.8

EXPERIMENTAL DATA

- NUMERICAL RESULT

0.7
0.6
0.5
0.4
~

03
02
0.1
0.0
-0.1
-02
-0.3
-0.6

-0.4

-0.2

0.0

0.2

0.4

0.6

Figure 3: Shock envelope of a hypersonic lifting body [40].


7
6

1 ,
2

Moo = 16.33

........
~HH~_,-------------45.0
30.0
37.5

ci+-----~~----~~~~~~~
0.0

7.5

15.0

22.5

Grid Pointa

Figure 4: Mach number distribution along the line of symmetry of a blunt body [56].

Chapter 4:
Point-Implicit Relaxation Strategies for Viscous,
Hypersonic Flows
P. A. Gnoffo
Aerothermodynamics Branch, NASA Langley Research
Center, Hampton, Virginia 23665, USA
ABSTRACT
An upwind-biased, point-implicit relaxation algorithm for obtaining the numerical
solution to the governing equations for three-dimensional, viscous, hypersonic flows
in chemical and thermal nonequilibrium is described. The algorithm is derived using
a finite-volume formulation in which the inviscid components of flux across cell walls
are described with a modified Roe's averaging and Harten's entropy fix with secondorder corrections based on Vee's Symmetric Total Variation Diminishing scheme.
Newton relaxation of the fully coupled equation set is employed on a cell to cell basis. Under-relaxation of the inviscid and over-relaxation of the viscous contributions
to the residual are implemented. Computational work is easily partitioned among
many processors in an asynchronous, dynamic mode for convergence acceleration.
An overview of the physical models employed herein for thermochemical none quilibrium is included. Several test cases and comparisons with experimental data are
presented involving hypersonic flow over blunt bodies which illustrate the qualitative
and quantitative capabilities of this approach.

NOMENCLATURE
Symbols
Boldface, lowercase symbols refer to vectors in parameter space. Boldface, uppercase
symbols refer to matrices in parameter space. An arrow over a lowercase symbol
refers to vectors in physical space, with x, y, and z coordinates.
a

A
B

frozen sound speed


Jacobian matrix of g with respect to q
Jacobian matrix of h with respect to q
mass fraction of species s
average molecular speed of molecule s

116 Computational Methods in Hypersonic Aerodynamics

Cv,tr

Gv,v

D.
es
eV
eV,s

f
g

h
h
hs
hv,s
H

I
kb,r
kj,r

m x , my, m z
ML
ML,INV
ML,vIS
ML,SRC

Ms
n

ii
n x , ny, n z
n
Nr

Ns
p
q
Qrad

rjINV
rjvIs

drag coefficient
heat transfer coefficient
reference heat transfer coefficient
lift coefficient
specific heat for translational-rotational energy
specific heat for vibrational-electronic energy
base diameter
effective diffusion coefficient for species s
energy per unit mass of species s
mixture vibrational-electronic energy per unit mass
vibrational-electronic energy per unit mass of species s
total energ~ per unit mass of mixture
flux vector in Cartesian space
inviscid component of flux vector relative to cell face
viscous component of flux vector relative to cell face
enthalpy or altitude
enthalpy per unit mass of species s
vibrational-electronic enthalpy per unit mass of species s
total enthalpy per unit mass of mixture
identity matrix
backward reaction rate coefficient for reaction T'
forward reaction rate coefficient for reaction r
unit vector tangent to computational cell wall
in X,y,z directions, respectively
components of
lift to drag ratio
unit vector tangent to computational cell wall
components of m in x,y,z directions, respectively
point-implicit Jacobian of flux terms
point-implicit Jacobian of inviscid terms
point-implicit Jacobian of viscous terms
point-implicit Jacobian of source terms
molecular weight of species s
iteration level index
unit vector normal to computational cell wall
components of ii in x,y,z directions, respectively
number density
number of reactions
number of species
pressure
vector of conserved variables
radiative energy transfer rate
right-hand-side residual vector
relaxation factor used with inviscid Jacobian matrices
relaxation factor used with viscous Jacobian matrices

Computational Methods in Hypersonic Aerodynamics 117

R
Rb,r
R"r
RN

Rs
R
s

t
T

Tv
u,v,w
U, V, W

X,Y,Z
Ys
a

universal gas constant


backward reaction rate for reaction r
forward reaction rate for reaction r
nose radius
shoulder radius
matrix of left eigenvectors of A
arc length
arc lengths in ii,~ m directions
approximation to gradient of characteristic variables
function of s
time
translational-rotational temperature
vibrational-electron-electronic excitation temperature
velocity components in X,y,z directions
velocity component in the ii, ~ and mdirections
Cartesian coordinates
mole fraction of species s
angle of attack relative to base plane normal
stoichiometric coefficient for reactants in reaction r

8p/8pE
stoichiometric coefficient for products in reaction r

8p/8Pr
eN
Ti
Tiv
()
(}e
(}R

A
~
A
/l
/lsi
~, Ti, (

P
Ps

a
T

parameters for defining minimum eigenvalue


eccentricity of nose
frozen thermal conductivity for translational-rotational energy
frozen thermal conductivity for vibrational-electronic energy
o for first-order, 1 for second-order approximation to inviscid flux
cone half angle in plane of symmetry
rake angle of base plane
eigenvalue of A
restricted eigenvalue of A
diagonal matrix of eigenvalues of A
mixture viscosity
reduced mass of species sand j
computational coordinates
mixture density
density of species s
cell face area
cross section for translational-vibrational energy exchange for species s
shear stress or relaxation time

8p/8p ev
w

dummy variable for ~, Ti, or (


vector of source terms
mass production rate of species s per unit volume

118 Computational Methods in Hypersonic Aerodynamics

Wv

vibrational-electronic energy source term


cell volume

Subscripts
e

i,j, k
I,J,/{
I
L
r

V
00

electron
indices of cell walls in the ~, TJ, ( directions
indices of cell centers in the ~,TJ, ( directions
dummy index for a cell wall
dummy index for a cell center
reaction r or species r
speCles s
vibrational-electronic
free stream

INTRODUCTION
The National Aeronautics and Space Administration's interest in viscous, hypersonic
flowfield simulation has grown in recent years in anticipation of the design needs
for space transportation and exploration over the next three decades, e.g. Walberg
[1]. Aeroassisted space transfer vehicles will use the upper layers of planetary atmospheres in hypersonic aerobraking maneuvers. The National Aero-Space Plane,
with its supersonic combustion ramjet engines, is being designed to achieve hypersonic speeds through the Earth's atmosphere in its climb to orbit. Various concepts
for a second-generation Space Shuttle are now being considered, as are concepts for
a space station crew emergency return vehicle. The external flowfield surrounding
such vehicles, as well as the internal flowfield through the scramjet engine, may be
significantly influenced by thermochemical nonequilibrium processes in the flow. Accurate simulations of these phenomena would provide designers valuable information
concerning the aerodynamic and aerothermodynamic character of these vehicles.
There are two major challenges to the simulation of flowfields in thermochemical
nonequilibrium around vehicles travelling at hypersonic velocities through the atmosphere. First, these simulations require modelling of the nonequilibrium processes in
the flow, frequently at energies where the models currently lack sufficient experimental or analytic validation. Nonequilibrium processes occur in a flow when the time
required for a process to accommodate itself to local conditions within some region
is of the same order as the transit time across the region. The equations and models
used in this chapter for non-equilibrium flow have been documented in Reference [2]
and were substantially derived from the work of Park [3,4] and Lee [5]. The conservation equations are formulated within the context of the LAURA algorithm (Langley
Aerothermodynamic Upwind Relaxation Algorithm) [7]. Calibration and validation
of the physical models intrinsic to this code, which must deal with nonequilibrium

Computational Methods in Hypersonic Aerodynamics 119


chemical and thermal processes, curve fits of thermodynamic and transport properties
at high temperatures, and radiative energy transfer, was first discussed in Reference
[8]. Other code development and calibration programs (e.g. Candler [9, 10], Park
and Yoon [11], Netterfield [12] Coquel et al. [13]) are now in progress within the area
of viscous, hypersonic, reacting gas flowfield simulations. Still, these calibration and
validation studies do not fully validate the flowfield simulations at velocities characteristic of vehicles returning to earth from the Moon or Mars. These simulations
should be used to help design ground-based and flight experiments for obtaining the
necessary data by performing parametric studies which identify critical aspects of the
physical models and locations in the flowfield which are most sensitive to unknowns
in these physical models.
Second, because of the large number of unknowns associated with chemical
species and energy modes and because of disparate time scales within the flowfield,
these simulations require algorithmic innovations to maintain numerical stability and
fully exploit supercomputer resources. Numerical stability is maintained through an
implicit treatment of the governing equations. A great variety of implicit treatments
is possible. For problems in which only the steady state solution is required, one is
free to evaluate any element of the difference stencil at any iteration (pseudo time)
level which facilitates the relaxation process. In the most rigorous implicit treatment,
all variables in all cells are simultaneously solved at an advanced iteration level, requiring the solution of a linearized equation set involving nI J K equations where n is
the number of unknowns at a point and I, J, and J{ are the number of computational
points in the three respective coordinate directions. The various forms of factored implicit schemes and line relaxation methods sequentially solve equation sets involving
nI, nJ, and/or nK variables. The point-implicit schemes, as utilized in the present
work, sequentially solve equation sets involving n simultaneous, linearized equations.
Further simplification is possible in chemical kinetic problems by linearizing contributions to the residual from only the source terms to alleviate problems of disparate
chemical time scales, resulting in methods which involve no matrix operations.
The essence of the point-implicit strategy is to treat the variables at the cell center
of interest implicitly at the advanced iteration level and to use the latest available
data from neighbor cells in defining the "left-hand-side" numerics. The success of this
approach is made possible by the robust stability characteristics of the underlying
upwind difference scheme. The basic algorithm requires only a single level of storage,
and numerical experiments show excellent stability characteristics, even when working
directly with the steady state equations (i.e., Newton iteration). Furthermore, the
algorithm is particularly amenable to generating solutions on unstructured grids [14]
and is efficiently implemented on vector or parallel processors [15]. Details of the
relaxation algorithm for the case of a gas in thermal and chemical nonequilibrium
are presented herein. Sample applications related to hypersonic flows over aero brakes
follow.

120 Computational Methods in Hypersonic Aerodynamics

FINITE-VOLUME FUNDAMENTALS
The integral form of the conservation laws applied to a single cell in the computational
domain is written

(1)
In Equation 1 the first term describes the time rate of change of conserved quantity
q in the control volume; the second term describes convective and dissipative flux
through the cell walls; and the third term accounts for sources or sinks of conserved
quantities within the control volume. The third term is identically zero for perfect-gas
flows but is required for flows in chemical or thermal nonequilibrium.
The finite-volume approximation to Equation 1 for a general, unstructured grid
is written

[ ~]
{)t L +

IIh

~f~
L...J m

n",(J"m

[.W H0] L

(2)

m=l

where {)q = qn+I - qn, {)t = t n+ 1 - tn, Jlh is the number of faces of cell L
having volume n, and subscript m refers to cell face m with surface area (J"m' The
quantity nm is a unit vector normal to cell face m in a direction facing away from
the cell center. The dependent variable q is defined at cell centers. The independent
variables x, y, and z are defined at cell corners.
The finite-volume approximation to Equation 1 for a rectangularly ordered, structured grid is written
{)q n
[----;5t ]J,J,J(

+ [ ;+1
+ [ ~+I
+ [ k+I

ni+1 (J"i+1

n'(J"']J
r
"'
t
, \

nj+I(J"j+I

nk+I (J"k+1

n(J"'
J J
nk(J"k

]1 ,\Y

(3)

]I,J = [w nh,J,I(

A shorthand notation for Equation 3 that will be used throughout this paper follows:

2:

[it+1

ii/+I (J"/+1

it . ii/(J"/]

[w n]L

(4)

/=i,j,k

Note in Equations 2-4 that uppercase integer variables I, J, f{, and L denote computational coordinates at cell centers, and lowercase integer variables i, j, k, 1, and m
denote cell faces or cell corners. For example, (J"i,J,I\ refers to the cell wall corresponding to indices 1-1/2, J, f{. (See Figure 1.) In the shorthand notation of Equation 4,
the integer variable I is used as a generic index for i, j, or k. This notation is convenient because most of the formulations for quantities at cell faces are independent
of the coordinate direction. The geometric quantities n, (J", and ii are easily derived
given the Cartesian coordinates of the cell corners. Details may be found in Appendix
A of Reference [6].
The formulations that follow are based on a rectangularly ordered, structured
grid. A first-order-accurate formulation of the inviscid equations on a structured grid

Computational Methods in Hypersonic Aerodynamics

cell center = (I,J)

j=3

cell corner = (i,j)

(I,j) or (i,J)

1
1
1

1
1
1

cell corner

.9~I~fM1te[. _
1

j=2

1
1
1
1
1

__

cell wall

1
1
1
____ ..l ____

1
1
1
- - - - -1- ___
1

I
I
I

I
I

I
I

ceiliwall

I
I
I
I

1
1

I
I

I
I

----r---I

1
1

_____ : ___ c.ll

121

~alL_~

____

--

--

- --

1
1
1

1
1
1

j=1

Figure 1: Cell indexing system with cell corners defined by lower case letters and cell
centers defined by upper case letters.
is identical to the formulation on an unstructured grid. The modifications required
to achieve second-order accuracy on an unstructured grid are not addressed in this
paper. However, it should be noted that the formulations for obtaining second-order
accuracy only involve modifications to the right-hand-side residual vector. The pointimplicit relaxation procedure that will be defined by the formulation of the left-handside matrix is independent of grid structure. Consequently, much of the development
which follows will carryover to unstructured grid formulations, as in the paper by
Thareja et al. [14].

CONSERVATION EQUATIONS
The inviscid, viscous, and source term contributions to the complete conservation
laws are considered separately for convenience. Let

(5)
where gl defines the inviscid terms and hi defines the viscous terms. The finite-volume
formulation of the conservation laws is now expressed as

L
I=i,j,k

[hl+ 1 0"1+1

h[O"l]

(6)

122 Computational Methods in Hypersonic Aerodynamics


In the case of a reacting gas flow in which thermal nonequilibrium is modeled using
a two-temperature approximation, the vectors q, g, h, and ware defined as

Ps
pu
pv
pw
pE
pev

(7)

PsU
pUu + pn,.
pUv + pny
pUw + pnz
pUH
pUev

(8)

- pDs OSn
~Y'
h
-

UTnx

VTny

WTnz
Tl

!tIY..

'IVas n

Tnx

Tny

Tnz

aT
!tIY..
"I aS n
"Iv aS n
I D ~
P ",N,
L....s=l ~ V,s s aS n

(9)
I D ~
P ",N,
L....s=l ~s s aS n

Ws

0
0
0

(10)

Qrad

Wv

The first element of the vectors defined in Equations 7-10 describes the species
conservation; the next three elements describe :1:, y, and z momentum conservation;
the fourth element describes total energy conservation; and the fifth element describes
vibrational-electronic energy conservation. The present model considers Ns = 11
species. Consequently, the vectors defined in Equations 7-10 are composed of a total
of 16 elements. Species 1 through 5 are the neutral components of air consisting of
N, 0, N 2 , O 2 , and NO. Species 6 through 10 are the ions corresponding to species 1
through 5, in which one electron has been removed. Species 11 are the free electrons.
Implicit in the use of a vibrational-electronic energy equation is the assumption that
the partition of energy in the vibrational, bound electronic, and free electron modes
among all species can be described by a single temperature, Tv. This approximation is based on rapid equilibration of vibra.tional and electronic energy and electron
translational modes [4, 16]. The translational and rotational energy modes of heavy
particles are assumed to be fully excitecl and described by temperature T.
The thermo-chemical nonequilibrium model is described in detail in Reference
[2]. Some specifics on their formulation is given in a later section but a brief overview

Computational Methods in Hypersonic Aerodynamics

123

is given below. The reactive source terms for the species conservation equations
are denoted by
The radiative energy transport term Qrad may be treated as
a source term in the energy equation though its effects are not considered in the
present analysis. Finally, the vibrational-electronic energy source term Wv accounts
for the mechanisms by which vibrational-electronic energy is lost or gained due to
collisions among particles in the cell. These mechanisms include the energy exchange
(relaxation) between vibrational and translational modes due to collisions within the
cell; the vibrational energy lost or gained due to molecular depletion (dissociation) or
production (recombination) in the cell; the electronic-translational energy exchange
due to elastic collisions between electrons and heavy particles; the energy loss due to
electron impact ionization; the rate of energy loss due to radiation caused by electronic transitions; and a term related to the work done on electrons by an electric field
induced by the electron pressure gradient minus the flow work due to electron pressure. The electron pressure flow work is normally considered as part of the electronic
enthalpy in the inviscid (convective) portion of the flux balance. Moving the electron
pressure from the convective term to the source term simplifies the expressions for
eigenvalues and eigenvectors of the Jacobian of the inviscid flux vector.

wS'

FORMULATION OF THE INVISCID TERMS


The inviscid flux vector at cell face I is defined

(11 )
where gL,1 = [fL . ndlNv.
The first term in braces is a second-order-accurate base approximation for gl.
The second term in braces provides the upwind-biased numerical dissipation. It is
a first-order dissipation when 0 equals zero. It is a second-order dissipation when ()
equals one.
The vector SI is defined

(12)
The matrix Ri 1 in Equation 12 and the matrices R/ and Al in Equation 11 are
related to the Jacobian of the inviscid flux vector g with respect to q in the following
manner:
A =
= R A R- 1
(13)

og

oq

The matrix R is the matrix of eigenvectors of A, and A is a diagonal matrix containing


the eigenvalues of A. The elements of these matrices, which are required at a cell
face, are evaluated as appropriate averages of quantities at adjacent cell centers. This
averaging procedure is discussed in the Appendix. The matrix IAI is a diagonal
matrix containing the absolute values of the eigenvalues of A with constraints on the

124 Computational Methods in Hypersonic Aerodynamics


minimum allowed magnitude of an eigenvalue given by

= {

(14)

The eigenvalue limiter ( was first used by Harten [17] to prevent the formation of
expansion shocks across a sonic line where one eigenvalue equals zero. Its application
is critical in blunt body flows to prevent instabilities, often in the form of reversed
flows at the stagnation point, which occur in the stagnation region where the repeated
eigenvalue U is near zero. The magnitude of (, which is nondimensionalized by
the free-stream velocity, is problem dependent. Vee [18] has suggested a functional
dependence of ( on the local values of sound speed and velocity. This relation has
been adapted for use in the present work as follows:

(15)
where (0 is a constant which generally varies from 0.01 to 0.4. Larger values of (0
are required for flows with extensive stagnation regions, as in the case of blunt body
flows.
The term (Xn)1 is a shorthand notation for Vx . f'i l and may be thought of as
the inverse of the projected distance between cell centers Land L - 1 in a direction
normal to cell face t. It is defined in Equations 79-81 in the Appendix. The variable
X is a generic computational coordinate running in the direction of increasing generic
index L.
The limiters used in the present work were introduced by Vee [19, 20]. They
involve symmetric functions of gradients in the neighborhood of the cell face, and
algorithms based on these limiters are refcITcd to as Symmetric Total Variation Diminishing (STVD) schemes. STVD schemes involve little extra programming work
over simple first-order algorithms because most of the quantities required in their
implementation are already available. These anti-diffusive corrections are defined as
(16)
where the min mod function returns the argument of smallest absolute magnitude
when all the arguments are of the same sign or returns 0 if the arguments are of
opposite sign. Equation 16 is evaluated element by clement of vector s.
The STVD limiter does not yield a strictly upwind biasing on the formulation
of the flux vector. Also, the scheme reduces to first-order at cell faces where there is
a sign change in the arguments of the min mod function. The present formulation
differs from that of Reference [20J in that the differences Ri 1 (qL - qL-d have been
scaled by (Xn k This scaling, which is eventually removed by the leading factor of
the numerical dissipation term in braces, reduces the effects of grid stretching on the
argument returned by the min mod function. The present treatment has been found
to be well suited to the range of numerical test problems considered herein.

Computational Methods in Hypersonic Aerodynamics

125

Equation 11 can be approximately linearized with respect to bqL in the following


manner. Define
(17)
where superscript n refers to the current value at cell center L, superscript n + 1 refers
to the new value to be computed at cell center L, and superscript * refers to the latest
available value at neighbor cell L - 1. The notation g~L refers to the inviscid flux
through cell face I evaluated using the latest available data from cell center L - 1 and
the predicted data at cell center L. Elements of the vector si 1im are also computed
using current data at cell centers Land L - 1. Substitute gZ,1 + AL,IbqL for gLr
in Equation 17 to obtain

where

lAd

RI

lAd

Ri 1 . In like manner, one can show that

The point-implicit discretization of the inviscid part of Equation 6 can now be


expressed by
(18)
I=i,j,k

I=i,j,k

[(AL,I+l

+ IA1+11)al+l

- (AL,I -IAd)atJ bqL

I=i,j,k

An application of Stokes theorem to the summation of AL,I and AL,I+l in Equation 18


will show that

I=i,j,k

I=i,j,k

Therefore, Equation 18 may be simplified as

l==i,j,k

l=i,J,k

where

(20)

126 Computational Methods in Hypersonic Aerodynamics

FORMULATION OF THE VISCOUS TERMS


The viscous stresses on a cell face with unit normal1i in the orthogonal directions ii,
~ and m are given by

Tnn
Tnl
Tnm

au
Al ( au + av + -aw) + 2J..LIaS n
aS n aS I aS m
( aS
av +
III
n ~~
aw+au)
J..LI ( aS
aS-m

(21 )

(22)
(23)

Sn, S[,
J. L

Sm
A

where U, V, and ltV are velocity components, and


and
are arc lengths
in the 1i, ~ and
directions, respectively. The variables and
are the viscosity
coefficients. All transport properties at cell face I are obtained as linear averages of
properties at adjacent cell centers.
The component of shear stress acting in the S direction ( S being a dummy
variable for x, y, or z ) on a cell face with unit norm Ii can be expressed

1n

(24)

Substitution of Equations 21-23 into Equation 24, collecting terms, and simplifying (see Reference [6]) yields the following relation for shear stress in the S direction:
=

as

as
as
au a~ au a1] au ac)
a1]1]n + ac Cn + a~ as + a1] as + ac as

J..LI(a~~n +

w . '\l~
~
w ~
w ~
AI(+ - . '\l1] + - . '\lOn.
a~
a1]
ac

(25)

where s is a dummy variable for u, v, or w corresponding to S = x, y, or z, and terms


like ~n or 1]1 are shorthand notations for "9~ . Ii or "91] . ~ respectively. A thin-layer
approximation in the X coordinate direction (X = ~, 1], or () simplifies Equation 25
by neglecting derivatives in the other two coordinate directions. Consequently,

Tns = IlI(

as 1 au
-a
+ --a
nshn
X
3 X

(26)

where n refers to the direction normal to a constant X surface, and the prime superscript refers to the thin-layer approximation. The use of Stokes relation,
=
and geometric identities have also been used in the simplification of Equation 25. The
viscous terms on the other two coordinate surfaces are also neglected in the thin-layer
approximation because their contribution to overall momentum and energy balance
is small. These approximations are valid so long as the boundary layer is relatively
thin and the X direction is approximately normal to the high gradient region.
Mass diffusion and energy conduction contributions to the viscous terms are
functions of gradients normal to the cell face. For example, the gradient of T in the

A -2J..L/3,

Computational Methods in Hypersonic Aerodynamics

127

normal direction n is expressed


aT = V(T)
a Sn

aT

,
a~ ",n

i'i

aT
a1] 1]11

aT
---rn

(27)

a('

The thin-layer approximation to Equation 27 is expressed


aT
=

aS

aT

---x
ox ~ n

(28)

Derivatives in the X direction are evaluated to second-order accuracy in computational space as follows:
au)*
( aX
I,L

UL

+ bttL -

au)*
*
n+l _
*
( OX
1+1,L = UL+1 - UL
- UL+ 1

Ur._l

ttL -

_
ttL -

(au)
OX 1+1

b
UL

Terms like ~~ (X = ~) are evaluated as follows assuming a rectangular ordering of


mesh points:
( au)
01] i+l,J,K =

*
41 ( UI,J+l,K
-

*
UI,J-l,/\

*
+ U!+l,J+1,K
-

*
)
U/+ 1 ,J-l,K

The derivatives in the directions along the face (i.e., those derivatives neglected in the
thin-layer approximation) have no functional dependence on the cell center. Therefore
the point-implicit treatment of the full Navier-Stokes equations is identical to the
thin-layer N avier-Stokes equations.
Now, define hi as a function of differences evaluated using currently available
data, for example (~~)I' and define hi,L as a function of differences using predicted
values at cell center L, for example (aaU)i
x ,L' These definitions permit the linearization
of the viscous terms to be expressed as follows:
hi,L

hi -

hi+l,L

h l +1

BI,L bqL

B I+ 1 ,L bqL

BI,L

_ ahi,L
aqL

_ Ol;i,L

B I+ 1 ,L

ahi+l,L
OqL

Ol;i+l,L
OqL

(29)

where

aqL

(30)

The point-implicit implementation of the viscous terms follows the example set
in the previous section on the inviscid terms.
I=i,j,k

I=i,j,k

128 Computational Methods in Hypersonic Aerodynamics


where
ML,vIS

[B/+ 1,LO"I+1

+ Bl,LO"d

(32)

l=i,j,k

In the case of the thin-layer Navier-Stokes equations, the summation would only include one of the i, j, or k directions, depending on the orientation of the computational
coordinates with the body.

FORMULATION OF THE SOURCE TERMS


Species Conservation
The mass rate of production of species s per unit volume is expressed
Nr

Ws = Ms L((3s,r - os,r)[Rf,r -

(33)

Rb,r]

,'=1

where N r is the number of reactions; Os,r and (3s,, are the stoichiometric coefficients
for reactants and products in the T' reaction, respectively; and Rf,r and Rb,r are the
forward and backward reaction rates for the T' reaction, respectively. These rates are
defined by

Rf,T
Rb,T

(34)

[ kf,T g(Ps/Ms)"s,r ]
=

[kb,r

(Ps / Ms ){3"r ]

where kf,T and kb,T are the forward and backward reaction rate coefficients, respectively, defined in [2], and Ns is the number of chemical species.
The reaction rate coefficients are explicit functions of T and Tv. Consequently,
the Jacobian of Ws with respect to q can be explicitly evaluated as follows:

(35)

where the differential relations between T and q and between Tv and q can be
expressed
N,

L(e s -

eV,s)dps

s=1

udpu

vdpv -

wdpw

dpE - dpev

(36)

N,

pCv,vdTv

dpev

L eV,sdps
s=1

(37)

Computational Methods in Hypersonic Aerodynamics

129

Total Energy Conservation


In preliminary work completed to date, the radiative energy transport term Qrad may
be treated in a purely explicit manner. Radiative energy transport has been calculated using the method of Hartung [21] based initially on converged, non-radiative,
nonequilibrium flowfield solutions. These radiative source terms are then held constant while the governing equations are relaxed again. In cases of strong radiation,
this relaxation process may require a slower introduction of the source through appropriate averaging of the old and new source terms. There is no point-implicit
contribution from this term in the algorithm.
Vibrational-Electronic Energy Conservation
The vibrational-electronic energy source term Wv may be subdivided into three functionally distinct sets of terms.

Wv

(Ri,r - Rb,r)t }

r:elec . mp.

L ps (ey,s - ev,s) +
{ s:mol.
< Ts >

{Qrad - Pe 9

3peR(T - Tv)

L ~~ }

s-te

(38)

.u }

The first set, the reactive source terms in the first pair of braces of Equation 38,
are composed of terms that are proportional to either Ws or to (Ri,r - Rb,r)' In the
first case, the proportionality factor, D., represents the average vibrational energy per
unit mass created or destroyed through recombination or dissociation of molecules.
In the simplest approximation, it is set equal to the average vibrational-electronic
energy, ev, though more comprehensive treatments which model preferential dissociation of vibrationally excited molecules may be employed. In the second case, the
proportionality factor, r , represents the average translational energy per mole lost
by a free electron in freeing another electron from a neutral heavy particle in reaction r through the process of electron impact ionization. It is approximated by the
ionization energy from an excited state of the target particle. Further detail on these
points is available in Reference [2] and the book by Park [16J. The point-implicit formulation of these terms treats the proportionality factor explicitly and the reaction
rates implicitly according to Equations 35-:37.
The second set, the relaxation terms in the second pair of braces in Equation 38,
model the energy exchange between heavy particle translational-rotational modes
and vibrational-electronic and electron translational modes. The first term in these
braces, which models the exchange between vibrational and heavy particle translational modes, can be approximated by

s:mol.

Ps

(ey,s - e~T,s)

< Ts >

T - Tv
pCv,v ----TV

(39)

130 Computational Methods in Hypersonic Aerodynamics


where

Ls=mol. PsI(Ms < Ts


( 40)
TV
Ls=mol. Psi Ats
The approximations in Equations 39-40 are made to reduce the number of thermodynamic and relaxation time variables to be carried through the calculation. Also,
direct evaluation of the equilibrium value, eir s' is more cumbersome than working
directly with the translational temperature T. This approximation degenerates as
the differences between T and Tl,T get very large, but it is believed to be consistent
within the total context of approximations made in the two-temperature model. The
vibrational relaxation time < Ts > is related to the correlations of Millikan and White
[22] through a number density weighting and a high temperature limiting correction
of Park [16].
"N.
~j=1,jf.e

~."
nJexp

[A s (T- 1/ 3 _ 0. 015 J.lsj1/4) "N.


~
~j=1,j#e nj

Tsp =

< Ts >

( - ~)-1
f7 sCsn

= Tt1W

18 . 4?~ ]

(41 )
(42)

T;

( 43)

Values of As in Equation 41 are 220 for N 2 , 129 for O 2 and 168 for NO and p is in units
of atmospheres. The two-temperature model should also have a corresponding term
relating the energy exchange of translational and electronic energy. This transfer has
not yet been formally included in the present work; however, the driving potential
is already based on both the vibrational and electronic energies and the relaxation
times are expected to be on the order of < Ts > so that the net effect on the present
model should be small. The second term in these braces in Equation 38 models the
direct exchange of translational energy between electrons and heavy particles. This
exchange rate is generally much slower than the previous term. Both terms in this
set are now proportional to the difference between the translational and vibrational
temperatures, T - Tv. Here again, the point-implicit formulation of these terms
treats the proportionality factor explicitly and the driving potential T - Tv implicitly
according to Equations 36-37.
The third set, the field-dependent terms in the third pair of braces of Equation 38,
are functions of properties at the cell center and at neighboring cells. (The first
two sets are functions only of properties at the cell center.) These terms include
radiative energy transport, work done by the electric field on electrons and electron
pressure flow work, combined into a single term. Radiative energy transport is treated
explicitly as described before. The other contribution to the field dependent terms
is also treated explicitly. In fact, in the cases tested to date with maximum electron
number densities approximately 4 percent of the total number density, omission of
this term has little effect on the flowfield.

Computational Methods in Hypersonic Aerodynamics

131

Point-Implicit Relaxation of Source Term


The source term in Equation 6 can be approximately linearized in the following manner:
(44)
where

OWL

ML,SRC

and the elements of

ML,SRC

= -OqL

(45)

are calculated as described above,

RELAXATION ALGORITHM
The governing relaxation equation is obtained by combining the results of Equations 6,20,32, and 45 and taking the limit as time step 8t goes to infinity. Thus,

(46)
where

ML

is the point-implicit Jacobian given by


( 47)

and r is the right-hand-side solution (residual) vector given by


rL

= -

[(gIH

+ hl+dal+l - (gl +

hl)ad

wLfk

(48)

I=i,j,k

Relaxation factors are used to control stability and convergence. Numerical tests in
Reference [15] indicate that underrelaxation is appropriate for the inviscid contribution to the residual, with r flNV > 1.5. Overrelaxation is appropriate for the viscous
contribution to the residual with I'fv IS > 0.5 provided relaxation sweeps are across
the boundary layer; otherwise, r fv IS ~ 1. The lower limits yield the fastest convergence rates but may lead to instabilities if the solution is far from convergence or if the
point-implicit Jacobian is "frozen", as discussed below, for too long. It is sometimes
necessary to chose r flNV ~ 3 and r fv IS ~ 2 to get past some difficult transients
in the early stages of the relaxation process that defy linear analysis. \Vhen these
transients pass it is then advisable to switch to the lower limits of these parameters
to get the best convergence rate. Convergence may eventually sta.ll at some point due
to limit cycles associated with the min mod function in Equation 16. This stalling
may be alleviated by again increasing the relaxation factors.
The solution vector rL and the Jacobian ML are evaluated using the latest available data. Consequently, the algorithm requires only a single level of storage. One
can solve for 8qL using Gauss elimination. Numerical experiments have shown that
pivoting is not required, and so the algorithm is easily vectorized. However, it is more
efficient to calculate and save Mr: 1 for large blocks of iterations (typically 20) and
solve for 8qL directly using
( 49)

132 Computational Methods in Hypersonic Aerodynamics


This algorithm requires no more work per computational cell than a purely explicit
formulation except for the effort needed to multiply a vector of residuals by a matrix.
For the case of three-dimensional flow with 11 species and two temperatures the
vector and matrix dimensions are (16 x 1) and (16 x 16). For the case of threedimensional flow of a perfect gas, the vector and matrix dimensions are (5x1) and
(5x5). In three-dimensional blunt body flow problems using approximately 60,000
cells, the "freezing" strategy was approximately 1. 7 times faster than the unmodified
algorithm for perfect-gas flow involving 5 unknowns per cell, and was approximately
2.8 times faster for the nonequilibrium case involving 16 unknowns per cell. The
solution vector is now updated according to
(50)
It should be noted that there is a large cost in computational memory required for the
Jacobian freezing. Preliminary tests indicate that the matrix ML may be replaced by
a diagonal matrix with elements related to the maximum eigenvalues of ML according
to the methods of References [11] and [23]. The consequences of this simplification
are still under investigation.
New values for T and Tv are obtained through a Newton-Raphson iteration based
on Equations 36 and 37. Thermodynamic properties and reaction rate coefficients
are advanced every iteration based on these updated values of T and Tv. Transport
properties are updated every four to twenty iterations.
The strategy used to drive the right-hand side of Equation 48 to zero should
take advantage of the host computer architecture and the physics of the problem.
Here, the solution is relaxed one plane at a time, and vector lengths are equal to
the number of cells in a plane as implemented on the CRAY 2. Numerical tests
indicate that relaxation sweeps which run from a wall across the boundary layer to
the opposite boundary and then back again are the most efficient for the blunt body
problem. Effects of a perturbation at a wall are felt at the opposite wall after one
sweep. Effects of a perturbation at one cell in a plane parallel to the wall require n
iterations to be felt by a cell whose index differs from the source cell by n.
The ordering of the sweeps may be used to speed convergence, but in numerical
tests performed to date, final, converged steady state solution is not effected. Thus,
one should be able to solve a large number of cells using a massively parallel processing
computer in which each cell (or small group of cells) is relaxed semi-independently
of its neighbor cells (cell groups) using its own processor. The expression "semiindependently" means that a cell (cell group) will need updated information from its
neighbor cells (cell groups), but neither the order that it receives this information nor
the lag time it takes for this information to arrive is critically important. As long as
each processor has immediate access to some level of information from its neighbors
(which could be stored locally), the execution stream could proceed uninterrupted
in a parallel, asynchronous mode. A crude simulation of asynchronous iteration,
discussed in Reference [24], demonstrated that computational cells could be advanced
in a random order without sacrifice of stability or convergence.
Asynchronous iteration has been tested on a four processor Cray II and eight pro-

Computational Methods in Hypersonic Aerodynamics

-0.0

-1.0

sum of six tasks


task 1

t:,.

task 2
task 3
task 4

task 5
task 6

-2.0

133

109,o(Error)

-3.0

- - - - - - - - _________s~n9Ie task

-4.0

-5.0

-6.0~~~~~~~~~~~~~~~~--~~~~~~-J

100

300

200

400

500

600

CPU time, s

Figure 2: Convergence histories for single-task and six-task, adaptive partitioned algorithms applied to problem of nonequilibrium, hypersonic flow over a blunt, axisymmetric body.
cessor Cray Y-MP in Reference [15]. In these tests, the flow domains were subdivided
into partitions with a single task assigned to each partition. Partition boundaries are
dynamically adjusted to concentrate relaxation sweeps in the regions that are slowest
to converge. Because no synchronization is required, all tasks (processors) may execute throughout the computation without interruption. A comparison of convergence
histories for the solution of hypersonic flow in thermochemical nonequilibrium over an
axisymmetric body is shown for a single task and a six-task, adaptive partition test
in Figure 2. The symbols show the error norm for each individual task of the six-task
run. The solid line shows the total error norm for the six-task run and the dotted
line shows the error norm for the single task run. Adaptive partitioning has allowed
the six-task case to converge to a lower error norm than the single task case for the
same amount of CPU time. Furthermore, the actual elapsed time for the six-task case
would be a factor of six smaller than for the single task case on a dedicated machine.

RESULTS AND DISCUSSION


Aeroassist Flight Experiment (AFE) - Wind Tunnel
The Aeroassist Flight Experiment (AFE) is a blunt, raked, elliptic cone designed to
obtain flight data in a hypersonic, nonequilibrium flow regime to test aerobrake design

134 Computational Methods in Hypersonic Aerodynamics


concepts and provide CFD code validation information [25]. The ground based wind
tunnel tests for this configuration do not simulate the nonequilibrium air chemistry
encountered in flight, but they do serve to check predictive technique and capabilities
on the actual vehicle shape, and so are quite valuable in the validation process.

Figure 3: Mach number contours in the plane of symmetry of the Aeroassist Flight
Experiment (AFE) model at Mach 10.
Mach number contours over the AFE model including the sting for Mach 10
flow and a = _5 are shown in Figure 3. The Reynolds number for this case is
159000 based on a model diameter of 9.322 em and laminar flow is assumed. Grid
adaption is used over the forebody to align the grid with the captured bow shock.
The contours clearly illustrate the sharp, captured bow shock, the captured shock on
the wind side of the cylindrical sting, and the free-shear layers emanating behind the
circular shoulder of the aerobrake.
Comparisons with experimental data of Micol [26] for pressure and heat transfer
are shown in Figure 4 on the forebody. Differences between numerical predictions
and experiments are generally very small and within experimental accuracy, except
for the heat transfer in the stagnation region. These differences are believed to be
caused by both the varying truncation error behavior as the coordinate singularity

Computational Methods in Hypersonic Aerodynamics 135

Experiment} C
----- LAURA
P

2.0

nOn

QY-~O

1.6
C

C ' _H_ 1.2


p

'Q....Q

0 Experiment} CH
- - LAURA
C-H, FR

f19-<l~ ~

\~

CH,FR

.8
.4

o~---~----~---~~-~

-.3

.3

.6

.9

s/L

Figure 4: Comparisons between prediction and experimental data for pressure and
heating over the AFE at Mach 10 and a = 0.

o
o

.8

q
q ref

88;
- O}E
.
180
xpenment

LAURA

.6
.4
.2
0

.5

1.0

1.5

s/L
Figure 5: Comparison between prediction (laminar) and experimental data for heating
on the sting supporting the AFE model at Mach 10 and a = 0.

136 Computational Methods in Hypersonic Aerodynamics


is approached from different directions and the predominance of near-zero-valued
eigenvalues in the upwind-biased algorithm. There is evidence that the problem is
exacerbated by extreme coordinate stretching in the direction normal to the body
[27].
Comparisons with experimental data of Wells [28] for heat transfer on the sting
are shown in Figure 5 for 0: = 0. The comparisons are within experimental accuracy (8%) on the wind side, except in the immediate vicinity of the peak heating
point where the free-shear layer impacts the sting. This comparison provides the first
verification of the predictive capabilities of the present algorithm in the hypersonic,
near wake of an aerobrake configuration.
Cylinder - Shock Tunnel
The flow around a 5.08-cm-diameter cylinder in high-temperature, pa,rtially dissociated nitrogen was computed, and fringe patterns are compared with the interferograms obtained by Hornung [29]. Free-stream conditions for this case are
Voo = 5.59 km/ s, Too = 1833 K, CN2 = 0.927, and CN = 0.073. The fringe
number F is determined by the expression

F =

t:.pL(l +
4160'\

CN)

(51)

where ,\ is the wavelength and L is the geometrical path, taken here as the cylinder
LAU RA, Park Model 1

M 00

Experiment, Hornung

=6.14
=12000

Re oo

T00 =1833K

Figure 6: Calculated and experimental fringe patterns over cylinder in partially dissociated nitrogen obtained in shock tunnel using chemical kinetic model of Park.
length equal to 0.1524 m (6 in). No attempt was made to account for changes in
the geometric path through the shock envelope due to three-dimensional effects at
the edge of the cylinder. There is reasonahly good agreement between the calculated

Computational Methods in Hypersonic Aerodynamics

137

and experimental fringe patterns in Figure 6, in which the chemical kinetic model is
due to Park as described in Reference [2]. A revision of Park's kinetic model was also
tested for this case in [8]. Some shifting in the fringe pattern was observed in the
calculation but no conclusion could be drawn as to which model best fit the data.
The calculated shock shape is in excellent agreement with the experimental data.
A thermal equilibrium, chemical nonequilibrium case was run by reducing the characteristic time for vibrational relaxation by a factor of 105 . The change in shock
standoff distance at the stagnation streamline was less than 5% as compared to the
thermochemical nonequilibrium case.
FIRE II Stagnation Point Heating - Flight Data
FIRE II flight data [30] consist of measured total heating rates (convective plus absorbed radiative), radiative intensity in the 0.2 to 6.2 eV range, and spectral radiation
in the 2.0 to 4.0 eV range. Only the stagnation point measurements are examined
in the present set of results. The vehicle was constructed with a layered heatshield,
consisting of three beryllium layers. The vehicle had a truncated, hemispherical shape
with a small corner radius. The calculations were performed at three points during
the early data period on the first heatshield, which is characterized by low radiation and significant thermochemical nonequilibrium. A fully catalytic wall boundary
condition is assumed. An equivalent sphere geometry was defined as in References
[31,32,33] with a nose radius of 0.747 m. The free-stream velocity equals 11.3 km/s,
and the altitude range covered by this early data period was between 84.6 and 67.05
km.

.
-01.0

Measured total
Measured total minus measured radiative
LAURA, Park model

~ Equilibrium VSL, Gupta

.8

.6

q ref
.4
.2

0
1630
Time, sec

Figure 7: Stagnation-point, convective heating predictions compared to experimental


data for the FIRE II flight test during early period in which radiative heating levels
were small and thermochemical noneqllilibrium effects were significant.

138 Computational Methods in Hypersonic Aerodynamics


The calculated convective heating rates from LAURA are compared with the
equilibrium, viscous shock layer results of [33] and with the experimental data in
Figure 7. The predictions are in excellent agreement with the total heating data
at the earliest data point, at which time there was no significant radiative heating. However, a fully catalytic wall boundary was assumed for this case, which may
be in contradiction to the possible formation of beryllium oxide on the surface. A
crude approximation to the actual convective heating rate is deduced from measured
quantities by subtracting the measured radiative heating in the 0.2 to 6.2 eV range
from the measured total heating. This approximation ignores radiation outside of
the measured range and does not account for the actual absorbtivity of the beryllium
heatshield. The inferred experimental convective heating is in good agreement with
the LAURA predictions, but this agreement is of dubious significance at the later
data period (t = 1637.5 s) because of the magnitude of the radiative contribution. Of
greater significance is the convergence of the equilibrium and nonequilibrium predictions at the lower altitudes, using two different approaches that provide independent
checks of the predicted convective heating.
Aerobrake Flight Simulations
The algorithm has been applied to simulate flowfields surrounding the Aeroassist
Flight Experiment (AFE) aerobrake and carrier and a proposed lunar return vehicle at selected trajectory points. More complete details of these simulations may be
found in Reference [34]. Both forebody and near wake flows are calculated. Vehicle
geometries are summarized in Table I. The trajectory points studied for each vehicle
are presented in Table II. The case labeled "afe-l" corresponds to one of the earliest
points on the AFE trajectory at which continuum analyses are credible in the shock
layer over the aerobrake. The case labeled "afe-4" corresponds to the peak heating
point on a nominal trajectory for return from geosynchronous Earth orbit. The case
labeled "lunar" corresponds to the peak heating point on a nominal trajectory for
return from the Moon. Computed aerodynamic coefficients for these cases at several angles of attack are recorded in Table III. An overview of the "lunar-a" case IS
provided below showing calculations over the forebody and near wake.
Table I - Vehicle
case
D, m R N , m
afe
4.2469 3.8608
lunar-a 7.3152 3.0480
lunar-b 7.3152 3.0480
lunar-c 13.7160 3.0480
Table
case VC<)' m/s
afe-l
9863.
afe-4
9326.
lunar
987l.

Parameters
Rs, m Oc
0.3861 60.
0.3048 70.
0.6096 70.
0.3048 70.

OR

eN

73.
90.
90.
90.

2.0
l.0
l.0
l.0

II - Trajectory Point Parameters


PC<), kg/m 3 TC<), K N ReooD N Re2D h, km
5.682E-06
189.
20,900
560 87.
4.293E-05
200. 143,000 5,600 75.
4.623E-05
210. 269,000 9,600
75.

----

---~---

Computational Methods in Hypersonic Aerodynamics 139

Table III
case
afe-1
afe-1
afe-4
afe-4
afe-4
lunar-a
lunar-a
lunar-b
lunar-c

- Aerodynamic Parameters
Q'
CD LID
CL
17. 0.366 1.37 0.267
22. 0.443 1.27 0.349
12. 0.302 1.43 0.211
17. 0.376 1.33 0.283
22. 0.429 1.22 0.352
O. 0.000 1.65 0.000
10. 0.232 1.53 0.152
10. 0.206 1.45 0.142
10. 0.239 1.58 0.151

No-slip and finite-rate wall catalytic boundary conditions [35] are used throughout. Aerobrake surface temperatures are held constant at 1750 K. Base surface
temperatures are held constant at 1500. Radiative energy transport is not included
in the computations. (Radiative heating accounts for about 5% of the total, maximum
heating for case "afe-4" and about 80% of the total heating for case "lunar-c".)
A new approach has been taken in the definition of the lunar aerobrake that
overcomes the problems associated with the axis singularity in the stagnation region.
This singularity may induce erratic behavior in the calculated heat transfer distribution. These problems were specifically addressed in [27] where it was noted that a
localized grid restructuring does much to overcome the problem. This concept was
extended to create a more uniform, singularity free surface mesh as shown in Figure 8.
A grid generation package for general blunt bodies with circular shoulders derived
from circular or elliptic cones has been coded to simplify the definition of singularity
free grids. As an added benefit, the new grid structure provides enhanced circumferential distribution of cells which feed directly into the wake. All of the lunar aerobrake
flows were generated with this new grid structure. Grid lines are constructed normal to the body. An algebraic stretching function is used to control distribution
of points across the shock layer. Mesh spacing at the wall is defined to enforce a
cell Reynolds number pallsl fL equal one. An adaptive grid routine aligns the outer
boundary with the bow shock so that the bow shock stands at 80% distance between
the wall boundary and the outer boundary of the computational domain.
Convective heating contours for case "lunar-a" at Q' = 10 are also presented in
Figure 8. Heating distributions for cases "lunar-a", "lunar-b", and "lunar-c" along
the plane of symmetry at Q' = 10 are compared in Figure 9. These distributions show
none of the erratic behavior evidenced in earlier work associated with the coordinate
singularity. The peak heating point for the "lunar-a" configuration occurs on the
shoulder for Q' = 10. The larger shoulder radius of the "lunar-b" configuration
reduces the peak shoulder heating (the reduction factor is slightly less than the square
root of shoulder radii ratio) at the expense of lower values for C L , CD, and LID.
The larger base radius of the "lunar-c" configuration raises the effective radius of
curvature of the stagnation point on this sonic corner body, thus reducing stagnation
point heating relative to the two smaller bodies, even though they all have the same
nose radius.

140 Computational Methods in Hypersonic Aerodynamics

Figure 8: Surface mesh and heating contours over aerobrake for case "lunar-a" at
a = 100.

60

lunar-a
lunar-b
lunar-c

50

40

'"E

-(.)

0-

30

'\
I, ' '-- --,
,,
,
I
,
k

20

,;,

/"

10

0
-20

10

x, ft

10

20

Figure 9: Comparisons of heating distributions for three configurations at a = 100.

Computational Methods in Hypersonic Aerodynamics

Figure 10: Grid in symmetry plane and two circumferential planes


behind lunar aerobrake "lunar-a" at Q: = 10.

III

141

near wake

The outer boundaries of the wake computational domain in the symmetry plane
are straight lines extrapolated from th~ last two points of the outer boundaries of
the forebody grid. The forebody grid has been aligned with the captured bow shock.
An exit plane location is specified at a fixed z location several base diameters behind
the base. The midpoint of the exit plane is calculated and a circle is defined with
circumferential discretization equivalent to the forebody exit plane. The outer shell
of the wake boundary is defined by straight lines between corresponding points on
the exit plane of the forebody and the circular exit plane of the wake. Analytically
defined stretching functions are used to control mesh size extending from the exit
plane of the fore body flow domain to the exit plane of the wake.
An inner control shell which defines the boundary between wake zone 2 (extending from the outflow boundary of zone lover the forebody) and wake zone 3
(extending from the aerobrake base) must also be defined. The intersection of this
inner control shell with the exit plane forms a circle with a radius approximating the
radius of the wake core (one quarter of the base diameter). The radial distribution of
points in zone 2 blends smoothly with the radial distribution of points in zone 1. The
radial distribution of points at the exit plane of zones 2 and 3 are algebraically defined
to enhance resolution of the wake core, where the free shear layers come together. The
radial distribution of points along the base of the aerobrake in zone 3 is algebraically
defined to help resolve the free shear layer in the vicinity of the boundaries of zones
2 and 3.

142 Computational Methods in Hypersonic Aerodynamics

(a) pressure

(b) streamlines

Figure 11: Pressure contours and streamlines in symmetry plane in near wake behind
lunar aerobrake "lunar-a" at a = 100. Contour levels defined by pi Poo F~ varying
between 0 and 0.04 with 81 levels, inclusive.
An initial flowfield is calculated to better define the positions of the free shear
layer and wake core. Grid control shells are then moved to enhance resolution of these
structures. The wake neck region, where the free shear layers come together and the
flow compresses and turns into the wake core through the action of the recompression
shock, may also be identified in the initial solution. The inner control shell is then
reconfigured into a funnel like shape in which the converging surface of the funnel
extends from the aerobrake shoulder to the wake neck and the tubular extension of
the funnel follows the wake core to the exit plane.
A representative grid is shown in Figure 10. Pressure contours in Figure 11 (a)
clearly show the recompression shocks surrounding the wake core, as defined by the
streamlines in Figure 11 (b) for case "lunar-a" at a = 100. The recompression shocks
are diffused somewhat because they are oblique to the grid aligned with the wake
core. Better resolution requires application of a solution adaptive grid strategy as in
reference [36). The translational and vibrational temperature contours, the subsonic
mach number contours and electron number densities in the near wake are recorded
in Figure 12. The subsonic flow domain in the wake is one measure of a relative
quiescent zone for payload packaging.
Caution must be excercised in the interpretation of near wake flow results for
continuum calculations at high a.ltitudes. For example, the computed Knudsen numbers in the base flow (i.e., the ratio of mean free path to local characteristic length)

Computational Methods in Hypersonic Aerodynamics

143

(a) translationalrotational temperature

(c) subsonic mach number

(d) electron number density


~

-~~--------

Figure 12: Contours in symmetry plane in near wake behind lunar aerobrake "lunara" at a = 10. Mach number contour levels vary between 0 and 1.0 with 11 levels,
inclusive.
for case "afe-4" are as large as 1.0 in the low-density region immediately behind the
shoulder. Knudsen numbers in the high-gradient bow shock are of order 0.10. The
local characteristic length is equal to Ipj\! pi and gives an approximation of the length
scale over which significant changes in the flowfield are accommodated. The continuum approximation breaks down when the Knudsen number is of order 1 or larger.
A calculation at a higher altitude earlier in the trajectory (case "afe-I") revealed
order 1 Knudsen numbers in the captured shock and a larger region of order 1 values
behind the corners of the aerobrake. Simulation errors due to the limitations of the
present equation set in these isolated areas of large Knudsen number require more
study, but may be more amenable to treatment by direct simulation methods [37] or
a modified Burnett equation analysis [38].

CONCLUDING REMARKS
An upwind-biased, point-implicit relaxation algorithm for obtaining the numerical solution to the governing equations for three-dimensional, viscous, hypersonic
flows in chemical and thermal nonequilibrium is described. The present thermochemical nonequilibrium model includes 11 species and two temperatures. The algorithm is derived using a finite-volume formulation in which the inviscid components

144 Computational Methods in Hypersonic Aerodynamics


of flux across cell walls are described with a modified Roe's averaging and Harten's
entropy fix with second-order corrections based on Vee's Symmetric Total Variation
Diminishing scheme. The relaxation strategy is well suited for computers employing
either vector or parallel architectures. It is also well suited to the numerical solution
of the governing equations on unstructured grids.
The essence of the point-implicit relaxation strategy is to treat the contribution of
cell-centered quantities to the right-hand-side residual of the cell center at an advanced
time step. The contributions of the neighbor cells to the right-hand-side residual are
taken at their current time step. The left-hand side Jacobian is constructed from
the point-implicit Jacobians of the inviscid, viscous, and source term contributions
to the right-hand side residual. The application of separate relaxation factors to the
inviscid and viscous contributions to the Jacobian can be used to enhance stability
and convergence rates. Newton relaxation of the steady equations is employed. The
inverse Jacobian from each computational cell may be saved over large blocks of
iterations for significant reduction in computer time per relaxation step at the cost of
increased memory requirements. Consequently, the algorithm works like an explicit
scheme with a single, additional step involving the multiplication of a matrix times a
vector, with dimension equal to the number of unknowns at a cell.
The relaxation process may be asynchronously implemented in adjacent subdomains. Sacrificing synchronization requirements keeps all processors active throughout the computation. The partition boundary of each process is dynamically adapted
to concentrate relaxation sweeps in the areas that are slowest to converge. This
strategy reduces total CPU time required to achieve a given error norm in several
test cases.
A code calibration program for the algorithm is then reviewed, with emphasis
directed toward hypersonic aerobrake applications. Comparisons between experimental data and numerical simulation focus on perfect-gas tests over a scale model
aerobrake and on flight and ground tests which challenge some aspect of the thermochemical nonequilibrium model. These comparisons are described as calibration
runs because they test elements of the numerical simulation, but no single data set
adequately simulates the full-scale aerobrake flight conditions. Perfect-gas flow predictions feature comparisons with experimental data for pressures and heating on an
aerobrake forebody and on the supporting sting. Agreement is generally very good.
Good agreement was also obtained with convective heating data from the FIRE II
flight experiment and with the fringe pattern data of Hornung obtained for Mach 6
flow of partially dissociated nitrogen over a cylinder. In both cases, themochemical
nonequilibrium models must be employed to obtain reasonable agreement with the
data; however, it is difficult to use these data to evaluate subtle differences in the
modeling of nonequilibrium phenomena.
Finally, a summary of results for aerobrake flowfield simulations at actual flight
conditions is presented. The solution methodology is highlighted on a simulation
of a hypersonic flow over a blunted, 70 degree cone at ten degrees angle of attack
near the peak heating trajectory point on return from the Moon. The methodology
features a new procedure for generating singularity free surface grids over blunt bodies.

Computational Methods in Hypersonic Aerodynamics 145


The simulations include the near wake flowfield which is important for describing the
aerothermalloads experienced by a payload attached to the base of a blunt aerobrake.

REFERENCES
[1] Walberg, Gerald D.: "A Survey of Aeroassisted Orbit Transfer" ,Journal of Spacecraft and Rockets, Vol. 22, No.1, January-February 1985, pp 3-18.
[2] Gnoffo, Peter A.; Gupta, Roop N.; and Shinn, Judy: "Conservation Equations and Physical Models for Hypersonic Air Flows in Thermal and Chemical
Nonequilibrium," NASA TP 2867 (1989).
[3] Park, Chul: "Problems of Rate Chemistry in the Flight Regimes of Aeroassisted
Orbital Transfer Vehicles", Thermal Design of Aeroassisted Orbital Transfer Vehicles, edited by H. F. Nelson, Progress in Astronautics and Aeronautics, Vol.
96, AIAA, New York, NY, 1985.
[4] Park, Chul: "Assessment of Two-Temperature Kinetic Model for Ionizing Air",
AIAA Paper 87-1574, June 1987.
[5] Lee, Jong-Hun: "Basic Governing Equations for the Flight Regimes of Aeroassisted Orbital Transfer Vehicles", Thermal Design of Aeroassisted Orbital Transfer Vehicles, edited by H. F. Nelson, Progress in Astronautics and Aeronautics,
Vol. 96, AIAA, New York, NY, 1985, pp 3-53.
[6J Gnoffo, Peter A.: "An Upwind-Biased, Point-Implicit Relaxation Algorithm for
Viscous, Compressible Perfect-Gas Flows," NASA TP 2953, February 1990.
[7J Gnoffo, Peter A., "Upwind-Biased, Point-Implicit Relaxation Strategies for Viscous, Hypersonic Flows," AIAA Paper 89-1972-CP (1989).
[8J Gnoffo, Peter A., "Code Calibration Program in Support of the Aeroassist Flight
Experiment," Journal of Spacecraft and Rockets, Vol. 27, No.2, pp 131-142,
March-April 1990.
[9J Candler, G.: "On the Computation of Shock Shapes in Nonequilibrium Hypersonic Flows," AIAA Paper 89-0312, January 1989.
[10] Candler, G. V. and MacCormack, R. 'vV.: "The Computation of Hypersonic
Ionized Flows in Chemical and Thermal Nonequilibrium", AIAA Paper 88-0511,
January 1988.
[11] Park, Chul and Yoon, Seokkwan: "A Fully-Coupled Implicit Method for ThermoChemical Nonequilibrium Air at Sub-Orbital Flight Speeds," AIAA Paper 891974, June 1989.

146 Computational Methods in Hypersonic Aerodynamics


[12] Netterfield, M. P.: "Hypersonic Aerothermodynamic Computations Using a
Point-Implicit TVD Method," 1st European Symposium on Aerothermodynamics for Space Vehicles, May 1991.
[13] Coquel, F., Joly, V., Marmignon C., and Flament, C.: "Numerical Simulation of
Thermochemical Non-Equilibrium Viscous Flows Around Reentry Bodies," 1st
European Symposium on Aerothermodynamics for Space Vehicles, May 1991.
[14] Thareja, R. R.; Stewart, J. R.; Hassan, 0.; Morgan, K.; Peraire, J.: "A Point
Implicit Unstructured Grid Solver for the Euler and Navier-Stokes Equations,"
AIAA-88-0036, January 1988.
[15] Gnoffo, Peter A.: "4synchronous, Macrotasked Relaxation Strategies for the
Solution of Viscous, Hypersonic Flows," AIAA Paper CP 91-1579, June 1991.
[16] Park, Chul: "Nonequilibl'ium Hypersonic Aerothermodynamics ," John \Viley &
Sons, Inc, 1990.
[17] Harten, Ami.: "High Resolution Schemes for Hyperbolic Conservation Laws,"
Journal of Computational Physics, Vol. 49 (1983), 3.57-393.
[18] Vee, H. C.; Klopfer, G. H.; and Montagne, J.-L.: "High-Resolution ShockCapturing Schemes for Inviscid and Viscous Hypersonic Flows," NASA TM
100097, April 1988.
[19] Vee, H. C.: "On Symmetric and Upwind TVD Schemes," NASA TM-86842,
September 1985.
[20] Vee, H. C.: "Construction of Explicit and Implicit Symmetric TVD Schemes
and Their Applications," Journal of Computational Physics, Vol. 68, 1987, pp.
151-179.
[21] Hartung, Lin C.: "Development of a Nonequilibrium Radiative Heating Prediction Method for Coupled Flowfield Solutions," AIAA Paper 91-1406, June 1991.
[22] Millikan, Roger C. and White, Donald R.: "Systematics of Vibrational Relaxation," J. Chem. Phys., vol 39, no. 12, Dec. 15, 1963, pp. 3209-3213.
[23] Eberhardt, Scott and Imlay, Scott: "A Diagonal Implicit Scheme for Computing
Flows with Finite Rate Chemistry," AIAA Paper 90-1577, June 1990.
[24] Gnoffo, Peter A.; McCandless, R. S.; and Vee, H. C.: "Enhancements to Program
LAURA for Computation of Three-Dimensional Hypersonic Flow", AIAA Paper
87-0280, January 1987.
[25] Jones, J. J.: "The Rationale for an Aeroassist Flight Experiment," AIAA Paper
87-1508 (1987).

Computational Methods in Hypersonic Aerodynamics 147

[26] Micol, John R.: "Experimental and Predicted Pressure and Heating Distributions for an Aeroassist Flight Experiment Vehicle in Air at Mach 10," AIAA
Paper 89-1731,1989.
[27] Grasso, Francesco and Gnoffo, Peter A.: "A Numerical Study of Hypersonic Stagnation Heat Transfer Predictions," Proceedings of the Eighth GAMM-Conference
on Numerical Methods in Fluid Mechanics, NNFM. Vol. 29, Vieweg, Braunschweig 1990, pp 179-188.
[28] Wells, William 1.: "Surface Flow and Heating Distributions on a Cylinder in
Near Wake of Aeroassist Flight Experiment (AFE) Configuration at Incidence
in Mach 10 Air," NASA Technical Paper 2954, 1990.
[29] Hornung, H. G.: "Non-equilibrium Dissociating Nitrogen Flow Over Spheres and
Circular Cylinders," J. Fluid Mechanics, Vol. 53, part 1 (1972), 149-176.
[30] Cauchon, D. 1.: "Radiative Heating Results from the Fire II Flight Experiment
at a Reentry Velocity of 11.4 Kilometers per Second," NASA TM X-1402 (1966).
[31] Sutton, K.: "Air Radiation Revisited," AIAA Paper 84-1733 1984.
[32] Zoby, E. V. and Sullivan, E. M.: "Effects of Corner Radius on Stagnation-Point
Velocity Gradients on Blunt, Axisymmetric Bodies," NASA TM X-1067 1965.
[33] Gupta, R. N.: "Navier-Stokes and Viscous Shock-Layer Solutions for Radiating
Hypersonic Flows," AIAA Paper 87-1576 1987.
[34] Gnoffo, Peter A., Price, Joseph M., and Braun, Robert D.: "On the Computation
of Near Wake, Aerobrake Flowfields," AIAA Paper 91-1371, June 1991.
[35] Stewart, David A.; Leiser, Daniel B.; Kolodziej, Paul; and Smith, Marnell:
"Thermal Respone of Integral Multicomponent Composite Thermal Protection
Systems," AIAA-85-1056 1985.
[36] Davies, C. and Venkatapathy, E.: "A Simplified Self-Adaptive Grid Method,
SAGE," NASA TM 102198, October 1989.
[37] Moss, James N.; Bird, Graeme A.; and Dogra, Virendra K.: "Nonequilibrium
Thermal Radiation for an Aeroassist Flight Experiment Vehicle", AlA A Paper
88-0081 1988.
[38] Moreau, S., Chapman, D. R., and MacCormack R. W.: "Effect of Rotational
Relaxation and Approximate Burnett Terms on Hypersonic Flow-Field Radiation
at High Altitudes," AIAA Paper 91-1702, June 1991.
[39] Roe, P. L.: "Approximate Riemann Solvers, Parameter Vectors, and Difference
Schemes," Journal of Computational Physics, Vol. 43 (1981), 357-372.

148 Computational Methods in Hypersonic Aerodynamics


[40] Vinokur, M.: "Flux Jacobian Matrices and Generalized Roe Average for an
Equilibrium Real Gas," NASA CR-177512, Dec 1988.
[41] Liu, Yen and Vinokur, Marcel: "Upwind Algorithms for General Thermochemical Nonequilibrium Flows," AIAA Paper 89-0201, January 1989.
[42] Grossman, B. and Cinnella P.: "The Development of Flux-Split Algorithms for
Flows with Non-Equilibrium Thermodynamics and Chemical Reactions," First
National Fluid Dynamics Conference, Paper 88-3596, July 1988, Cincinnati,
Ohio.

APPENDIX
Definition of A, R, R -1, and A
The Jacobian of g with respect to q is expressed

U(o" - c.)
irnx - Uu
irny - Uv
1'rnz-Uw
irU - UH
-Uev

c&n X

un x (1 - (3) + U
-f3uny + vnx
-f3un z + wnx
-f3uU + Hnx
evnx

c.ny
-f3vn x + un y
vny(l - (3) + U
-f3vn z + wny
-f3vU + Hny
evny

c$n Z

-f3wn x + un z
-f3wny + vn z
wnz(l- (3) + U
-f3wU + Hnz
evnz

0
f3nx
f3ny
f3nz
f3U +U
0

0
</Jnx
</J n y
</Jnz
</JU
U

(52)

The similarity transformation matrices Rand R -1 are defined

bsr / a 2
u/a 2
v/a 2
w/a 2

0
Ix
Iy
Iz

iJ(u 2 +v 2 +w 2 )-'Yr
iJa 2

a 2bsr - c/Yr

-V

R- 1

-w
ir - Ua
ir + Ua
-evir

0
cs /2a 2
mx (u + an x )/2a 2
my (v + any)/2a 2
m z (w + an z )/2a 2
W (H + aU)/2a 2
0
ev/ 2a2

f3uc s
Ix
mx
-f3u + anx
-f3u - anx
f3 ue v

f3vc s
Iy
my
-f3v + any
-f3v - any
f3 vev

0
cs /2a 2
(u - an x )/2a 2
0
(v - any)/2a 2
0
(w - an z )/2a 2
0
(H - aU)/2a 2 -/ f3a 2
ev /2a 2
1/a 2
f3wc s
lz
mz
-f3w + an z
-f3w - an z
f3 we v

f3c s
0
0
f3
f3
-f3e v

(53)

-c s
0
0

a 2 - ev

(54)

The diagonal matrix of eigenvalues of A is defined by

U 0 0
0
0
0
0 U 0
0
0
0
0
0
0
0 0 U
0
0 0 0 U +a
0
U -a 0
0 0 0
0
0
0 0 0
0
U

(55)

Computational Methods in Hypersonic Aerodynamics

149

The variable c. is the mass fraction of species s.


(56)

c. = P./p

The variables (3, , and


respect to q.

ir

are related to the partial derivatives of pressure with

(57)

Pe
= -R- -

(58)

8p _
8pE 8p
8p ev

R
pr
pCv,tr r=l,re Mr
pCv,v Me

8p
RTq
u2 + v2
+
(3
2
8ps
M.
The variable a is the frozen speed of sound.

is =

+ w2

N.

a2

I>.is + (3[H - u

v2

w2]

ev

(3e. - ev,s

(59)

= (1 + (3)'E..

=1

(60)

This definition of a 2 comes from the evaluation of the eigenvalues of A. The variable
R is the universal gas constant, and M. is the molecular weight of species s.
The variables n x , ny, and n z are the x, y, and z components of a unit vector
normal to a computational cell face, and U is the normal component of velocity
through the cell face, defined by
U = un x

vny

(61 )

wn z

The two unit vectors I and are defined such that ii, ~ and are mutually orthogonal
(i.e., nil; = nimi = limi = 0). The velocity components in the land
directions,
tangent to the cell face, are then defined by

(62)

W = um x

vmy

W1n z

(63)

In the matrices defined above, the first row and column correspond to the N.
species continuity equations. Subscript s refers to row s and species s, and subscript
r refers to column r and species r where both sand r vary from 1 to 11 in the present
model. Note in Equation 59 above that Tq equals Tv when s is an electron; otherwise,
Tq equals T. Further details of the derivations may be found in Reference [2].
The variables at cell faces are evaluated as follows:
C1 ai
C1

is,/

R-T. +
= ----'!1
Ms

+ aLl

+ 1
2+ 2+ 2

(31 U I

v21

wI

- (3l e s,l - lev,s,1

(64)

(65)

150 Computational Methods in Hypersonic Aerodynamics

(66)
(67)

where d is a dummy variable representing u, v,


weighting parameters are defined by

tv, C s ,

C1 =

( TL )
TL- 1

C2 =

( PL )
PL-l

If, ev, (3, </;, e., and ev,s' The

(68)

(69)

This averaging procedure is a departure from the Roe's method [39] used in
the perfect gas version of this algorithm [6]. It evolved over the early code development period using a series of numerical tests which revealed a sensitive dependence
of the TVD (non-oscillatory) shock capturing capabilities on the definition of al and
1's,/. The weighting on sound speed and temperature is motivated by the observation that the thermodynamic state of the gas on the free-stream (low temperature)
side of a strong shock is of negligible importance in the consideration of momentum
and energy balance across the shock. The weighting on the other variables follows
the example of Roe. The present formulation neither enforces Roe's property U on
the pressure jump between two adjacent cells nor does it provide an average sound
speed al which exactly satisfies Equation 60 as a function of cell averaged variables.
However, it has provided robust convergence characteristics in the vicinity of strong
shocks (Moo > 30) without significant smearing of the shock transition zone using
a very simple formulation. Vinokur [40], Liu and Vinokur [41], and Grossman and
Cinnella [42] have proposed averaging schemes which enforce Roe's property U. These
formulations have not been investigated here.
Geometrical Relations
The recommended, second-order-accurate formulations for face-centered metrics, which
are required in the evaluation of the viscous dissipation terms across cell faces, are
presented below.

V~i,J,J( =

n I ,J,Idifi - 1 ,J,I';

+ ifi,J,Id + n l - 1 ,J,Idifi ,J,J( + ifi+1,J,K)

(70)

4n l ,J,J(n l - 1 ,J,I(
nl,J,J(( ifi ,J-l,J(

+ ifi+ 1 ,J-l,Id + nl,J-l,K( ifi,J,J( + ifi + 1 ,J,J()

(71)

4fh,J,J(nl,J-l,K
n I ,J,K(ifi ,J,K-l

+ ai+I,J,J(-d + nl,J,J(-I(a i ,J,I( + ai+l,J,K)


4n l ,J,J(n I ,J,J( -I

nl,J,K( if l-l,i,K

+ if l-l,i+1,1.) + n l - 1 ,J,J( (al,j,K + if l,j+1.K)


40, I ,J,J( 0, I -1 ,J,I(

(72)
(73)

Computational Methods in Hypersonic Aerodynamics

+ ih,j,Id + D[,J-l,KUh,j,K + ih,j+1,Id

DI,J,K(ihj-l,K

'9TJ[,j,K

151

(74)

4DI,J,[{D[,J-l,K
DI,J,K(ih,j,K-l
'9 TJ[,J,k

+ ih,j+1,[{-d + Dl,J,[{-1(8[,j,[{ + 8[,j+1,Id

(75)

4D[,J,KD[,J,K-l

'9 (i,J,K

D[,J,K(8[-1,J,k

+ 8[-1,J,k+1) + D[-1,J,K(8[,J,k + 8[,J,k+d

(76)

4D[,J,KD[-1,J,K
D[,J,K(8[,J-l,k

'9([,j,K

+ 8[,J-l,k+l) + D[,J-l,K(8l,J,k + 8l,J,k+d

(77)

4D I ,J,I,D[,J-l,K

'9 (I,J,k

D I ,J,K(8[,J,k-l

+ 8[,J,k) + DI,J,K-l(8I,J,k + 8I,J,k+d

(78)

4D[,J,I{D[,J,K -1

In the case of the thin-layer Navier-Stokes equations, only the vectors defined
by Equations 70,74,78 are required, depending on the orientation of the coordinate
system, The dot product of these vectors with the corresponding unit normal to the
cell face (recall, xn = '9X . ii) can be approximated as follows,
~ni,J,K

TJn[,j,K

DI,J,[{( i7i-l,J,K

i7i,J,K)

D[-I,J,g(i7j,J,K

.i7 i+1,J,K)

(79)

4D[,J,KD I - l ,J,K

DI,J,K (i7I,j-l,K

i7[,j,K)

D[,J-l,K( i7 [,j,K

i7[,j+1,K)

(80)

4D[,J,KD I ,J-l,K
(nI,J,k

D[,J,K(i7I,J,k-l

i7[,J,k)

D[,J,K -1 (i7I,J,k

i7[,J,k+1)

(81)

4Dl,J,KD[,J,K-l

Another useful formulation in the programming of the thin-layer Navier-Stokes


equations involves a geometric relation between the unit normal to a cell face and the
gradient of the computational coordinate which defines the cell face.

( ~: )
( ~~ )
(

~~)

(~ii ns)i,J,K

(82)

i,J,K

(83)
I,j,K

(84)
I,J,k

where s is a dummy variable for x, y, or z.

@No copyright is asserted in the United States under Title 17, U.S. Code. The U.S. Government
has a royalty-free license to exercise all rights under the copyright claimed herein for Governmental
purposes. All other rights are reserved by the copyright owner.

Chapter 5:
Flux-Split Algorithms for Hypersonic Flows
P. Cinnella (*), B. Grossman (**)
(*) ERC for Computational Field Simulation, Mississippi
State University, P.O. Box 6176, Mississippi State,
MS 39762, USA
(**) Department of Aerospace and Ocean Engineering,
Virginia Polytechnic Institute and State University,
215 Randolph Hall, Blacksburg, VA 24061, USA
ABSTRACT
This study reviews numerical techniques for the simulation of fluid flows spanning the range of reactive regimes, from local chemical equilibrium to full thermochemical non-equilibrium. In particular, characteristic-based algorithms are considered, of the flux-vector and the flux-difference type, which are becoming ever more
popular due to their accurate rendition of physically complicated, shock-wave dominated flows. Consideration is given to the problems associated with modeling the
thermo-chemical behaviour of reactive mixtures of thermally perfect gases, and a
few topics of current research are outlined.
1. INTRODUCTION
In recent years, a renewed interest in the field of hypersonic aerodynamics has
been registered among the scientific community. Vehicles like the Space Shuttle
or the European Hermes are fully operational or in an advanced design stage, respectively. Their conception would not have been possible without some knowledge
of chemical and thermal non-equilibrium phenomena and how they affect the performance of engines, lifting surfaces and materials, to name a few examples. The
Aeroassisted Orbital Transfer Vehicle (AOTV), conceived as an essential tool for
possible exploratory missions on Mars and for efficient satellite management in high
earth-centered orbits!, is anticipated to make use of aerobraking in conditions that
are not reproducible on earth in existing experimental facilities, nor will be in the
near future. Other aircraft, like the National Aero-Space Plane (NASP) or the European Hypersonic Transport Vehicle, are becoming national priorities and subjects
for commercial competition. In this context, the availability of accurate, robust
and efficient simulation tools for the study of the complex physical phenomena that

154 Computational Methods in Hypersonic Aerodynamics

occur in the hypersonic regime becomes an essential priority for the scientist and
the designer2. To this end, several computer codes have been recently developed 3 - 8
which are capable of representing geometrically complex three-dimensional flowfields
with non-equilibrium chemistry and thermodynamics. Most of them use upwind
technology in the discretization of the non-linear convective terms that appear III
the governing equations.
The essential core of this work is the analysis of the numerical techniques that
are at the center of the current, upwind-based simulation capabilities. In particular,
some classes of flux-split algorithms will be discussed, with the intent of showing
how these methodologies are applicable to the study of reactive flows. Originally developed for perfect gases (e.g., see the survey papers of Harten, Lax and Van Leer 9 ,
and Roe IO ), these techniques have produced accurate solutions of shock-wave dominated flows. Their essential feature is the treatment of the convective terms in the
Navier-Stokes equations using a characteristic-based differencing. In particular, the
three most popular types of approaches will be discussed, namely the flux-vector
splittings of Steger and Warming ll and Van Leer l2 , and the flux-difference splitting
of Roe l3 . Extensions of these algorithms to flows in local chemical equilibrium l4 - 2o
and to finite-rate chemistry flowS 21 - 27 have appeared in the literature, including in
some instances consideration of thermodynamic non-equilibrium effects.
The subject matter of this work will be organized in the following way: the influence of high-temperature, high-velocity effects on the thermodynamic behaviour
of mixtures of gaseous species will be examined in 2, and the influence on their
chemical behaviour will be discussed in 3. The governing differential equations for
viscous, heat-conducting flows which are out of chemical and thermal equilibrium
will be presented in 4. The reduced form valid for flows in local chemical equilibrium will also appear. Non-equilibrium thermodynamic and transport phenomena
will be briefly summarized in 5. Flux-split algorithms for the treatment of chemically and thermally active flows will be detailed in 6, and efficient techniques for
time-accurate or steady-state calculations will be mentioned in 7. Finally, some
conclusions will be drawn in 8.
2. THERMODYNAMIC MODELS
Hypersonic flow is characterized by regions of high temperatures, induced by
the strong shocks associated with the high velocities involved. In this context,
deviations from the ideal gas behaviour are warranted, and manifest themselves in
the activation of internal energy modes which behave non-linearly with temperature,
in the onset of chemical reactions, and ultimately in the establishment of local
thermodynamic non-equilibrium.

Computational Methods in Hypersonic Aerodynamics

155

2.1 Internal Energy


In recent years, a great deal of effort has been spent on physically accurate
modeling of "real gas" effects, which is the somewhat misleading terminology used
to indicate departures from the low temperature ideal gas laws. A useful review
of the current literature is presented by Gnoffo28. A major controversy among researchers has revolved around the proper representation of species internal energy,
and in particular the appropriate derivation of the former from the results of statistical mechanics. Specifically, the species internal energy is derived from the partition
function, and although there is no argument on the fact that the translational energy is independent of the internal modes of a particle (atom or molecule), there
is no general agreement on whether some or even all of the internal modes can be
considered to be independent, and what is the "proper" partition of modes into
independent groups, if at all possible. The concept of independence of contributions to the partition function is essential, because independent modes give rise to
summable contributions to the internal energy.
A general representation for the species internal energy per unit mass, particularly suitable for the modeling of problems where thermodynamic non-equilibrium
plays a key role, is the following

(1)
In the above, eAT) is the contribution due to translation and internal modes that
can be assumed to be in equilibrium at the heavy-particle translational temperature T. Moreover, en, (Tn,) is the contribution due to internal modes that are in
thermodynamic non-equilibrium, meaning that they are not in equilibrium at the
translational temperature T, but may be assumed to satisfy a Boltzmann distribution at a different temperature Tn,. Finally, ee, (Te) is the contribution from
the electronic states in a particle, or from free electrons when present, where it
has been recognized that the "electronic cloud" will reach internal equilibrium very
rapidly24, hence only one electron-electronic temperature Te has been used for all
of the species. Cases when the electronic contributions are considered coupled with
the vibrational and rotational states can be handled by this representation if the
non-equilibrium temperature Tn, corresponds to these coupled modes, and the term
ee, is dropped.
Equation 1 is valid for a very broad range of physical conditions, and is susceptible to simplification and reduction to simpler models when the flow regime is less
"extreme". Typically, the contribution in equilibrium at the translational temperature es will represent translational and rotational modes, the non-equilibrium term
en, will coincide with vibrational energy, and the electron/electronic contributions
will be either neglected 29 , or considered to be in equilibrium at the electronic temperature Te 28. Some authors suggest that partitioning vibrational and rotational
modes is not appropriate for hypersonic flow conditions 24 ,3o, and in that case en,
would represent the combined rotational/vibrational contribution. Other practical

156 Computational Methods in Hypersonic Aerodynamics


models that have been proposed 31 consider only one combined vibrational/electronic
temperature, which is tantamount to the assumption that a common Tn exists and
has to be used for both en. and ee" for all of the species s.
In the following, a gas mixture composed of N species will be considered, and
three major representations of the thermodynamic state of the mixture will be studied. The first one will be denoted Full Non-Equilibrium Themodynamic Model,
whereby all of the species are allowed to have terms such as es , in equilibrium with
the heavy-particle translational temperature T, and such as ee" in equilibrium with
the electron/electronic temperature Te. Moreover, the first M species will contain a
non-equilibrium contribution en,. Free electrons will be assigned to the last index,
N, and will have only a translational contribution to the internal energy, eeN(Te).
The presence of two different translational temperatures in the former model,
one associated with heavy particles and the other describing the state of free electrons, creates many complications in the derivation of governing equations 32 and basic gasdynamic properties 33 . Consequently, a Simplified N on- Equilibrium Thermodynam{c Model will be considered in the following, whereby the electron/electronic
contributions are not modeled by means of their own temperature, but are lumped
together with either es or en,. This model has been shown to produce accurate
results for physically challenging problems 8 , when the efficiency with which the different particles can reach a common translational temperature is not a significant
issue. The definition of internal energy per unit mass, Equation 1, is modified by
simply dropping the third term on the right hand side.
A further simplification of the thermodynamic model may be achieved for flows
where there is enough time for the internal modes to reach equilibrium at the translational temperature T, typically when supersonic to low hypersonic regimes are
considered. This Equilibrium Thermodynamic Model will feature only one contribution to the internal energy in Equation 1, namely es . The usual ideal gas model
is a special case of the Equilibrium Thermodynamic Model, when vibrational and
electronic modes are neglected, and translational and fully excited rotational contributions are included in es .

It is convenient to express
volume

es , en"

and ee, in terms of specific heats at constant

ee, =

Te
cVe,,(T)dT,

(2a, b, c)
where h f, is the heat of formation of species s at the reference temperature T ref .
The mixture value for the internal energy per unit mass, e, may be written using a
standard mass-fraction averaged summation, as follows
N

s=1

s=1

s=1

s=1

(3)

Computational Methods in Hypersonic Aerodynamics

157

where the species mass fraction C s has been introduced, defined as the ratio of
species density Ps to mixture density p = L:~ Ps. Similarly, the frozen specific
heats at constant volume, cv , and at constant pressure, cp , may be obtained by
means of the mixture rule, as follows
N*

N*

Cv =

2::::: csc v ,

N*

2::::: cscp , = 2::::: cs(cv , + Rs) ,

cp =

s=l

s=l

(4a, b)

8=1

where N* equals N -1 for the Full Non-Equilibrium Thermodynamic Model, and


N for the other two models, and Ri is the species gas constant. The mixture gas
constant R, taking into account only heavy particles when the Full Non-Equilibrium
Thermodynamic Model is utilized, is written as follows
N*

R=

2::::: C

R8 =

p -

Cv

(5)

8=1

A very important subcase of the Equilibrium Thermodynamic Model is local


chemical equilibrium, whereby it is assumed that the kinetic rates in the flowfield
are fast enough to bring the chemical reactions that occur in the system virtually
instantaneously to their equilibrium value, dictated by the Laws of Mass Action.
This assumption has played a significant role in the modeling of real gas effects due
to the simplification that it brings in the analysis, as will be shown in subsequent
sections. It may be useful to point out that if the flowfield is considered to be in
local chemical equilibrium, then it follows from thermodynamic considerations that
only two state variables are necessary to describe its condition at a given time for
every point in space. In the following, density and temperature will be selected
as the two fundamental state variables for this subcase 34 . The equilibrium specific
heats can be obtained as functions of temperature and density derivatives of the
mass fractions, as follows

(6a, b)

where the mixture enthalpy h = e + RT has been introduced. Unless otherwise


noted, throughout this chapter a partial derivative with respect to the temperature
will be taken at constant density, and vice versa.

158 Computational Methods in Hypersonic Aerodynamics

2.2 Thermal and Caloric Equations of State


Most flows of practical interest in the hypersonic regime will involve low to
moderate pressures and densities. In these instances, the gas particles will be weakly
interacting, that is their intermolecular forces will be negligible after a characteristic
distance of the order of the particle diameter. This assumption is widely utilized
in the scientific literature dealing with hypersonic gasdynamics, and one of its consequences is that each species in a mixture will behave as a thermally perfect gas,
whose internal energy is not function of density (or specific volume). Mixtures of
thermally perfect gases were considered in the previous subsection.
Another result of importance, stemming from the previous one, is the validity
of Dalton's Law, whereby the mixture pressure is equal to the sum of species partial
pressures. The Thermal Equation of State, which is Dalton's Law expressed as a
relationship between the mixture pressure and the translational temperatures, will
read for the Full Non-Equilibrium Thermodynamic Model
N*

P=

L PsRsT + PNRNTe = pRT + PNRNTe,

(7)

s=l

where the mixture gas constant R has been defined in Equation 5. The Simplified Non-Equilibrium Thermodynamic Model and the Equilibrium Thermodynamic
Model will utilize a similar Thermal Equation of State, but the contribution from
the electron/electronic temperature Te will be dropped, due to the fact that free
electrons, when present, will have the same translational temperature as the heavy
particles.
Unlike the ideal gas case, the state relationship of the pressure to the specific internal energy cannot be written directly in general, but occurs implicitly
through the translational temperatures. For the Full Non-Equilibrium Thermodynamic Model, Caloric Equations of State such as Equations 2c and 3 will relate
internal energy, or portions thereof, to temperature. Moreover, when the knowledge of the non-equilibrium temperatures Tn, is required, additional relationships
such as Equation 2b will have to be utilized. Iterative procedures are necessary for
the determination of the relevant temperatures from known values of the internal
energy contributions, due to the inherently non-linear character of these equations,
but they are found to be numerically efficient in most circumstances 27 . Once T and
Te are found, the pressure is evaluated from Equation 7.
The Simplified Non-Equilibrium Thermodynamic Model and the Equilibrium
Thermodynamic Model will still necessitate a caloric equation of state such as
Equation 3, which in general will be non-linear in the translational temperature
T. Practical models have been proposed 27 where the non-linearity is removed, and
the temperature can be eliminated between Equations 3 and 7 to yield an Equation
of State linking pressure to internal energy directly. However, barring special cases

Computational Methods in Hypersonic Aerodynamics 159


such as the above-mentioned, the simplicity of the ideal gas equation of state is lost
in hypersonic reacting flows.

2.3 Speeds of Sound


The speed of sound is a thermodynamic property that plays a key role in most
numerical algorithms for the simulation of high speed flows. An ideal gas will have
a unique, well defined speed of sound. However, when the gas mixture is thermochemically active, at least two different speeds of sound may be encountered. The
frozen speed of sound is defined by freezing chemical composition and thermodynamic non-equilibrium phenomena at the local (instantaneous) values, whereas the
equilibrium speed of sound assumes that composition is in local equilibrium and the
internal energy contributions have reached equilibrium at the local temperature.
Unfortunately, the previous definitions are not completely satisfactory when
used in conjunction with the Full Non-Equilibrium Thermodynamic Model, due to
the fact that two translational temperatures are present. Specifically, the thermodynamic temperature is usually identified with the temperature that can be defined from the thermal velocity mean-square value 35 , and when there are more
than one of these temperatures the link between macroscopic and microscopic definitions becomes ill-defined. Usually, the heavy-particle temperature is considered
to be the thermodynamic value, and the free-electron temperature is treated as a
non-equilibrium contribution 33 . Results for the frozen speed of sound obtained by
means of the previous assumption have appeared in the scientific literature 33 , but
the problem has not yet been completely settled.
In general, the choice of the speed that enters the description of the propagation
of information is dictated by the chemistry model, to be discussed in 3. When flows
that can be studied using either the Simplified Non-Equilibrium Thermodynamic
Model or the Equilibrium Thermodynamic Model are considered, the frozen speed
of sound has been found to be the relevant quantity27, as long as the gas mixture is
not considered to be in local chemical equilibrium. When the latter assumption is
utilized, then the equilibrium speed of sound is the key player. Formally, the frozen
speed of sound for the Simplified Non-Equilibrium Thermodynamic Model can be
written as follows
2

a ==

(8ap )
P

= 'Y_RT = 'Y_(p)
8,p,/p,e n ,

(8)

where s is the entropy, and l' is the ratio of frozen specific heats l' = cp/c v . The
Equilibrium Thermodynamic Model will feature the same result, however the nonequilibrium internal energy contributions en, will formally disappear from the previous expression. The interestingly simple final formula for the frozen speed of sound
was described by Clarke and McChesney36.
The special sub case of the Equilibrium Thermodynamic Model that features
flows in local chemical equilibrium will have an equilibrium speed of sound defined

160 Computational Methods in Hypersonic Aerodynamics

as
a

==

Cap7P )

= rRT = r(~),

P
where the isentropic index r = r(p ,T) is given by the following
s

(9)

P '"' a[RsT - (r -l)e 1


=,- + --RT
p
s
~ ac

s=l

,_ --

ah/ aT - 1 +
ae/aT -

if + T I:~=1 Rs( Oc s / aT)

(lOa, b)

---=-"-="-----'----'---'C
.

In the above, l' is the ratio of the partial derivative of enthalpy with respect to
temperature at constant density to the partial derivative of the internal energy with
respect to temperature, where the latter is the specific heat at constant volume C v .
The use of the isentropic index preserves the simplicity of the formula for the sound
speed, but the functional dependence of this index from density and temperature is
relatively involved.
There are at least five different "gammas" that can be defined for a gas in
chemical equilibrium: the isentropic index r, the ratio of specific heats, = cp/c v ,
the ratio of frozen specific heats l' = cp/c v , the ratio of enthalpy and internal energy
derivatives 1', and the ratio of enthalpy to internal energy
= h/ e = 1 + p/ pe. For
a frozen flow, the first four of these quantities will reduce to the same value, but for
general flow conditions their values will remain different.

,*

2.4 Practical Thermodynamic Models


The formulation detailed in the previous subsections is very general, and allows
the utilization of virtually any practical thermodynamic model for the determination of the specific functional form of internal energy, specific heats, and related
properties. In the current scientific literature, a few models have emerged as the
ones that are preferred by most researchers for numerical simulations of high-speed,
non-equilibrium flowfields. In particular, two non-equilibrium and two equilibrium
models will be described in the following, due to their broad utilization for practical
calculations.
The first non-equilibrium model considers translational and rotational modes
to be fully excited and in equilibrium at a common temperature, whereas the vibrational modes are in non-equilibrium. The electronic excitation, when not neglected
altogether, is considered to be in equilibrium either with the vibrational contributions or with free electrons at the common electron/electronic temperature Te. In
the context of the Full Non-Equilibrium Thermodynamic Model, this first practical
model will specialize the expression for the species internal energy, Equation 1, to
the following

es =nsRsT + h~
3
eN =2RNTe,

+ en. (Tn,) + ee.(Te) ,

s = 1, ... ,N*(= N-1),

(lla,b)

- - - - _ .__. _ - - -

Computational Methods in Hypersonic Aerodynamics

161

where the reference temperature has been set to the absolute zero, and h~ is the
corresponding species heat of formation, which is zero for free electrons. The coefficient ns will be equal to 3/2 for a monatomic gas, it will be 5/2 for a diatomic or a
linear polyatomic molecule, and will equal 6/2 for a nonlinear polyatomic molecule.
The relationships such as Equations 2b and 2c that link vibrational and electronic
contributions to their respective temperatures are usually specialized using simple
harmonic oscillator formulas 35 for the vibrational modes and only the first few electronic states
e

- R
n, -

Ne"

'\""

s ~
i=l

8.
V,S,!

(12a,b)

e e .' ITn,,_ 1 '


v,~,t

where Ne,s is the number of characteristic vibrational temperatures 8 v ,s,i for species
s, and the sum over electron energy levels whose degeneracy factors are gi and
characteristic temperatures are 8i is typically truncated after the first or second
contribution. In the context of the Simplified Non-Equilibrium Thermodynamic
Model, the previous relationships will have to be modified to make Te disappear

s=l, ... ,N*(=N),

(13)

where the non-equilibrium contribution en, will contain both vibrational and electron terms

and free electrons, when present, will be in equilibrium at the (only) translational
temperature T.
The second practical non-equilibrium model is somewhat simpler than the first
one. It assumes that the vibrational and electronic modes will exchange energy very
efficiently, and consequently they will reach a common vibrational/electron temperature Tn valid for all species that possess an equilibrium contribution. There has
been some indication that the previous assumption is valid for air during atmospheric reentry37. This model is less computationally expensive than the previous
one, because only one non-equilibrium energy equation will be necessary instead of
either M or M + 1. Moreover, it may be used again in both Full and Simplified
contexts, provided that the previous Equations 11-14 are modified by substituting
Tn for every entry that read Te or Tn,.
The first practical equilibrium model is a direct simplification of the nonequilibrium models considered, in that it considers translational and rotational contributions to the species internal energy at their fully excited values, whereas the
vibrational modes are included by means of a simple harmonic oscillator formula

162 Computational Methods in Hypersonic Aerodynamics


Again, only the first one or two electronic levels are usually considered.
In the second model, fourth-order polynomials 38 have been used to curve fit
C v ,. Many coupling and non-harmonic effects are introduced in these curve fits, but
the coefficients are valid in a range of temperatures, usually up to 6,000 K, and
this could turn out to be a serious limitation to their practical use. However, the
methodology is not affected by these practical considerations, and improved curve
fits are becoming available 39 , which extend its range of validity up to 30,000 K. It is
noteworthy that these curve fits have been derived with thermodynamic equilibrium
in mind, but they have been used for non-equilibrium calculations 28 by simply subtracting translational and rotational contributions and interpreting what is left as
a curve fit of coupled vibrational and electronic modes, in equilibrium at a common
temperature Tn. Consequently, the curve-fit approach becomes an alternative way
of modeling problems with only one non-equilibrium temperature, as seen previously
in conjunction with the second practical non-equilibrium model.
Although rotational modes have been assumed to be in equilibrium at the translational temperature when all of the above practical models have been discussed, it
is a well known fact that rotational non-equilibrium plays a key role in the study of
shock-wave structures 40 . On the other hand, inside a shock wave vibrational contributions are frozen 32 , and consequently an expression such as Equation 1 can still be
used for the species internal energy, where en, represents rotational contributions,
and the electronic modes are also negligible 41 .
3. CHEMISTRY MODELS
High-speed, high-temperature flows are typically characterized by some form
of chemical activity. A classification of chemical phenomena according to the time
available for the completion of reactions is usually employed for order-of-magnitude
estimates of the flowfield properties (and results in the definition of Damkohler
numbers 2 ). A typical fluid-dynamic time TFD, of the order of some relevant length
divided by an average flow speed, can be defined for the flow under consideration.
Then a typical reaction time TCH,j can be estimated, generally the time necessary
for halving or doubling the quantity of a certain species for a given temperature and
initial composition, when only reaction j is active. Three cases can occur

TFD

~TCH,j'
~ TCH,j,

(16)

~ TCH,j'

In the first case, reaction j has no time to occur, and a frozen flow can be assumed with respect to that specific reaction. The second case is the general case of
finite-rate chemistry whereby the actual kinetics of the reaction must be taken into
consideration. In the third case, reaction j has a virtually infinite time to evolve,
and consequently it will reach its equilibrium value given by the Law of Mass Action.

Computational Methods in Hypersonic Aerodynamics

163

In practice, calculations are performed where all of the chemical reactions are
supposed to be in equilibrium, frozen, or evolving with a finite rate. Perfect gas
results are a particular class of frozen chemistry simulations. Moreover, from the
above discussion it should be clear that frozen and equilibrium models are important limiting cases of the real-life finite-rate chemistry situation, which requires
simulating the actual kinetic behaviour of the chemical system.
3.1 Finite-Rate Chemistry
A general simulation of chemical effects can be achieved for a system containing
N species where J reactions take place

' ,]X 1
VI'

'X-->."X
fIX N
' 2 + ... + VN'
+ V'2 X
N ~ VI' 1 + V2/IX
' 2 + ... + VN'
,]

,]

,],J

j = 1, ... ,J,

,}'

(17)
where the v:,j and the V~,j are the stoichiometric coefficients of the reactants and
products of species Xs in the j-th reaction, respectively. Then, the rate of production
of the s-th species, w s , may be written as a summation over the reactions, as
follows 35

ws

==

dPs
dt = Ms

g(

",
[N
- Vs,j) kt,j

~(VS,j

PI )
MI

v; .
,J

- kb,j

gN (

v!'.]

PI )
MI

,J

'

s =l, ... ,N,

(18)
where M s is the species molecular mass. For reaction j the forward and backward
reaction rates, k t,j and kb,j, are assumed to be known functions of temperature,
and are related by thermodynamics
(19)
where Ke,j is the equilibrium constant, which is a known function of the thermodynamic state. Using Equation 19, Equation 18 can be rewritten as follows
J
_ dps
",
Ws = dt = Ms '2)Vs,j - Vs,j)kb,j
)=1

[N
_

lie,j II (MJ -IIN


1=1

PI

ViI,J.

PI
(MI)

vOl
I,J

1=1

s =1, ... ,N,

(20)
where the term in the brackets goes to zero when equilibrium compositions are
reached for ps, because it reduces to a formulation of the Law of Mass Action

j = 1, ... ,J,
valid for equilibrium concentrations.

(21)

164 Computational Methods in Hypersonic Aerodynamics

The finite-rate chemistry model described remains valid in conjunction with all
of the thermodynamic models discussed in the previous section. However, the presence of thermodynamic non-equilibrium in the flowfield is likely to exercise some
effect upon the reaction rates, especially when diatomic or polyatomic molecules
are involved, whose vibrational modes may be excited. Several attempts to model
the interaction between chemical and thermodynamic non-equilibrium have been
recorded 42 - 44 . In most cases, the forward and backward reaction rates are redefined so that they are functions of both translational and non-equilibrium temperature(s). The simplest model employs an "effective" temperature defined as the
geometric average of translational and vibrational temperatures, to be used for the
molecular dissociation reactions 37 . More refined models try to model "preferential
dissociation" of the molecules with excited vibrational modes 43 , although there is
only controversial evidence that this is a first-order effect for most flows of interest 45 .
Park 46 suggests that the relationship between forward and backward rates through
the equilibrium constant, Equation 19, may be incorrect when effective temperatures enter the picture, which would invalidate the standard procedure of measuring
and/ or calculating only one of the two sets of reaction rates and using the equilibrium
constant to recover the other set. The presence of an electron/electronic temperature Te further complicates the picture, due to the appearance of more effective
temperatures 46 .
The previous chemistry model is valid for gaseous systems and involves homogeneous reactions only, therefore excluding solid or liquid surface as well as photochemical reactions. At the present time, wall effects in general and wall catalyticity
in particular are still not entirely understood, let alone modeled. Wall treatment in
the hypersonic regime is typically accomplished by assuming either a fully catalytic
solid boundary, and consequently chemical equilibrium at the local temperature and
density, or a frozen wall, with corresponding zero gradients of mass fractions. A few
attempts of modeling finite-rate catalyticity in terms of absorption and desorption
phenomena have been initiated 47 , but an accurate simulation of solid boundaries
for high-speed flows has yet to come.
3.2 Chemical Equilibrium
There are many problems that render a finite-rate computation into a nontrivial task. The most important one is the uncertainty in the reaction paths and
in the reaction rates 2 ,35. Although the composition of a given mixture in chemical
equilibrium as a function of thermodynamic variables can be determined quite accurately by theoretical and experimental means, there is very scarce knowledge of
the actual mechanism with which a given reaction occurs. Many paths, leading to
intermediate products or highly unstable excited states, are available for virtually
every reaction. It is a delicate theoretical work to rule out some of the possibilities
with respect to others, and it is an impressive experimental work to confirm these
results in order to obtain kinetic rate data for specific reaction paths. Moreover, the
numerical simulation of flows in chemical non-equilibrium requires the inclusion of

Computational Methods in Hypersonic Aerodynamics

165

N species continuity equations into the algorithms instead of only one global mass
conservation equation, as will be discussed in 4. Also, the finite-rate equations
become extremely stiff when the flow is close to chemical equilibrium48 . All of
the above reasons render a numerical simulation based upon chemical equilibrium
very appealing from a computational standpoint, when compared with finite-rate
calculations. In addition, for a large number of flows of practical interest results of
considerable accuracy can be obtained using this simplifying assumption.
From Equation 20, it is easy to recognize the limiting cases of frozen and equilibrium flows. When the backward rate kb,j - t for any j, then the species rates of
production go to zero, and it can be shown that the species mass fractions are unchanged if no diffusion occurs. On the other hand, when the backward rate kb,j - t 00
for any j, then the terms in brackets will tend to zero, because the equilibrium composition is maintained, and the species rates of production reach a limit value that
will balance convection, diffusion and unsteady terms. Consequently, local chemical
equilibrium corresponds to the limiting value of reaction rates going to infinity at
every point in the flowfield 34 . This limiting value will be reached while keeping two
state variables constant, which will allow for a unique description of the equilibrium
state. A close examination of the species continuity equation and of the combined
First and Second Laws of Thermodynamics indicates that local chemical equilibrium
calculation should be performed at constant density and internal energy49. At this
stage, the unknowns to be determined are species densities and temperature.

The equations to be solved for the determination of the N species densities


and the temperature will be Ne mass conservation equations written for an equivalent number of elemental species, N - Ne equations corresponding to a possible
formulation of the Laws of Mass Action, and one energy equation of the type given
in Equation 3, specialized for thermodynamic equilibrium 34 . Several methods for
the solution of these non-linear equations have been utilized, and a relatively recent review may be found in Liu and Vinokur 50 . Roughly, two different classes
of algorithms have been developed, one based upon optimization methods, namely
minimization of free energy and/or maximization of entropy, and the other based
upon the direct solution of the Laws of Mass Action, possibly with the introduction
of techniques for reducing the number of unknowns, therefore reducing number of
equations and computational time 49 .
Probably the most important gas mixture in engineering is air at standard
conditions, which corresponds to 79% nitrogen and 21 % oxygen in volume, with
traces of other gases. Solutions of the equilibrium composition problem for vast
ranges of density and internal energy have been obtained, tabulated and utilized for
the determination of thermodynamic properties of the mixture in terms of curvefit values 50 ,51. The importance of these curve fits should not be underestimated,
but unfortunately all the results obtained will be valid for the specific composition
studied only. Even slight changes in the base mixture would require new tabulations.

166 Computational Methods in Hypersonic Aerodynamics

Consequently, aside from problems involving the fundamental base air mixture, the
simulation of flowfields in local chemical equilibrium for a general gas mixture has
to rely on the efficient coupling of chemical composition solvers, based upon the
previously mentioned governing equations, with flowfield solvers, for the "standard"
Euler or Navier-Stokes equations.
3.3 Thermodynamic Properties for Flows in Local Chemical Equilibrium
After the local equilibrium composition at a given density and internal energy
has been determined, the different thermodynamic state variables can be evaluated.
Among them, several have a significant practical interest, for example speed of sound
a, given by Equation 9, isentropic index r, given by Equation lOa, and pressure, as
well as transport properties such as viscosity coefficient and thermal conductivity.
The evaluation of all of these quantities is made possible by the knowledge of species
densities and temperature at equilibrium, but it also requires the values of the
partial derivatives of the mass fractions with respect to temperature and density, as
it appears from Equations 10. In general, these derivatives are determined from the
governing equations by means of the Implicit Function Theorem 34 , due to the fact
that Laws of Mass Action and Mass Constraints are the relationships that implicitly
define the behaviour of mass fractions with changing temperature and density.
For the specific important example of a standard air mixture, tabulations and
curve fits are available for a broad range of density and temperature for virtually all
of the thermodynamic and transport properties of interest 50 ,51. Again, for a generic
mixture of thermally perfect gases either specific tabulations have to be created
prior to the solution of a flow problem, or the chemical equilibrium solver has to be
coupled with the flow solver.
3.4 Practical Chemistry Models
It was previously mentioned that a detailed knowledge of the kinetic behaviour
of a chemically reacting gas mixture is necessary when finite-rate chemistry investigations are deemed necessary. In particular, the specific reaction paths leading to
dissociation, ionization, and atomic rearrangement are required. In recent years, a
lot of attention to the kinetics of air and of hydrogen/ air mixtures has been registered in the scientific community, spurred by the drive towards hypersonic flight and
the necessary propulsive tools to achieve it 2. A compilation of the most recent data
for the kinetics of air may be found in Park 37 , where attention is given to thermodynamic non-equilibrium and its influence upon the reaction rates through effective
temperature, as previously discussed. Detailed studies of hydrogen/air mixtures
have been published recently52.
When local chemical equilibrium calculations are performed, the actual reaction
paths are no longer necessary. Chemical equilibrium is dictated by thermodynamics,
namely the entropy has to be a maximum for a system at constant density and
internal energy35. Even when chemical reactions are actually necessary in order to

-- -

----------

Computational Methods in Hypersonic Aerodynamics

167

formulate the Laws of Mass Action, any set of independent reactions can be used
to obtain the same final result. The treatment of chemical reactions for equilibrium
chemistry becomes a bookkeeping device.
4. GOVERNING EQUATIONS
The governing equations for flows in thermo-chemical non-equilibrium represent the conservation and/or time variation of physical quantities such as mass of
a species in the mixture, momentum, non-equilibrium energy contributions, and
energy, established either for a control volume of arbitrary size (finite or infinitesimal), or for a generic point in the flowfield. The former approach results into an
integro-differential form,
, the latter into a non-linear differential system.
4.1 Integral Form-Thermo-Chemical Non-Equilibrium
The governing equations written in integral form are valid both in regions of
smooth flows and across discontinuities 53 , where they can be shown to reduce to
the Rankine-Hugoniot jump conditions in the limit of infinitesimal size. The control volume employed for the derivation is allowed to move and deform with time,
although many important applications take advantage of the simpler case of fixed
control volumes. For an arbitrary volume V, closed by a boundary S, the governing
equations in integral form read

~t JJfv QdV + t(S -

Sv)' ndS =

JJfv

WdV,

(22)

where Q is the vector of conserved variables, W is the vector of source terms, and
Sand Sv are the inviscid and viscous flux vectors, respectively. The unit vector n
is normal to the infinitesimal area dS and points outwards. In conjunction with the
Simplified Non-Equilibrium Thermodynamic Model, and utilizing a Cartesian frame
of reference (Xl, X2, X3) whose unit vectors are i, j, and k, respectively, the vectors
Q and W read
PI
P2

Q=

PN
PUI
PU2
PU3
Ple nl
PMe nM
pea

WI

W2
WN

W=

2::. P.g'
2::. P.g'
2::. P.g'

l
2
3

Ql
QM

(2::. P.g.) . u

(23a, b)

168 Computational Methods in Hypersonic Aerodynamics

and the flux vectors S, Sv read

S=

Pl(U-UV)
pz(u - uv)

-PI V I
-PZ V2

PN(U-UV)

-PNVN
Tlli + TIZj + T13k
T21 i + TZ2j + T23 k
T31 i + T3zi + T33k
-P1enl VI - qnl

pu 1 (u - U V) + pi
pU2(U-UV)+pj
pU3(U - uv) + pk
P1enl(u - uv)

(24a,b)

PMenM(u - UV)
peo(u - UV) + pu
where

e =(TllUl + TZIU2 + T31U3)i + (TI2 Ul + T22U2 + T32 U3)j+


(TI3 Ul

+ T23 UZ + T33 U3)k -

q - qR -

j=1

s=1

2..= qnj - 2..= Pshsvs .

(25)

In the previous formulas, the mass-averaged velocity U for the mixture has been
utilized, defined in terms of species velocities Us as follows
N

pu =

2..= psu s .

(26)

s=1
The difference between species and mass-averaged velocities IS a specIes diffusion
velocity Vs
.5 = 1, ... ,N.
(27)
Vs=us-u,
The velocity Uv that appears in the inviscid flux vector is the control surface velocity,
which is identically zero when the control volume is fixed with time. The first N
equations are species continuity equations, relating the time rate of change of species
densities Ps to convective and diffusive transport and to the creation/destruction of
the species due to chemical reactions. The species rate of production Ws has been
defined by Equation 18, when finite-rate chemistry models were discussed. Following
species continuity are the three components of the momentum equation. The source
term :Z::::s Psgs represents body forces, e.g. gravity and electric fields, and the viscous
fluxes involve the components of the shear stress tensor, Tij. After momentum, M
non-equilibrium energy equations are written, describing the creation and evolution
in time and space of the non-equilibrium components of the internal energy. The
viscous flux associated with these equations describes the heat transfer due to the
transport of non-equilibrium energy by diffusion, pse n , Vs, as well as the conductive

Computational Methods in Hypersonic Aerodynamics

169

heat transfer, qns. The source terms for the non-equilibrium energy equations, Qs,
will be described in more detail in 5. Finally, the global energy equation describes
the time evolution of the total internal energy per unit volume pea = pe + pu6/2.
The viscous flux in this case represents viscous dissipation, convective heat transfer
due to both equilibrium and non-equilibrium portions of the internal energy, q
and 2:: j qnj respectively, radiative heat transfer, qR, and diffusive heat transfer,
2::s Pshsvs, where hs is the species enthalpy, hs = e s + RsT.
In addition to the equations presented, thermal and caloric equations of state
will relate pressure and temperatures to equilibrium and non-equilibrium internal
energy components, as discussed in 2.2, and the viscous flux vector entries will
be expressed in terms of the entries in the vector of conserved variables, as will be
described in 4.4 and 5.5.
The vectors that appear in the governing equations for cases when the Full N onEquilibrium Thermodynamic Model is employed are very similar to those already
discussed, but they contain one additional equation, namely a non-equilibrium energy equation for free electrons 33 . At the present time, numerous assumptions and
approximations are used in the derivation and several simplifications are made to
this equation 28 . The reader is referred to the work by Park 37 , and Grossman, Cinnella, and Eppard 32 for some examples of the use of this equation in hypersonic
flows. The whole problem of multiple translational temperatures and their effect
upon the gasdynamics of non-equilibrium flows is still an object of current research.
Simplifications in the thermophysical model employed for the analysis of nonequilibrium-flows will bring about corresponding changes in the governing equations.
When only one common vibrational/electronic temperature appears in the thermodynamic model, only one non-equilibrium energy equation is solved instead of M,
which can be written by summing up the contributions from each non-equilibrium
component of the internal energy. In this instance, if free electrons are present,
their kinetic energy is usually neglected 21 , and the complications arising from the
presence of two translational temperatures are ignored 37 . Similarly, the Equilibrium
Thermodynamic Model can be handled by simply dropping all of the non-equilibrium
energy equations, along with the non-equilibrium contributions to the global energy
equation. The whole process is equivalent to setting M equal to zero. Inviscid flows
are modeled by dropping the viscous flux vector, in conjunction with any of the
thermodynamic models discussed.
4.2 Integral Form-Local Chemical Equilibrium
The previous forms of the governing equations are written for finite-rate chemistry problems. When the important special sub case of the Equilibrium Thermodynamic Model that deals with local chemical equilibrium is considered, the vectors
that appear in Equation 22 are further simplified. The global continuity replaces the
N species continuity equations, the non-equilibrium energy contributions disappear,

170 Computational Methods in Hypersonic Aerodynamics

and the final form of Q and W reads

PUp 1 )
PU 2

(28a, b)

( PU 3
peo
where only body forces are present in the vector of source terms W. Inviscid and
viscous fluxes read

s=

p(u - uv)
)
PU1(U - uv) + pi
pU2(U-UV)+pj ,
( PU3(U - uv) + pk
peo(u - uv) + pu

(29a, b)

where

e =( Tn U1 + T21 U2 + T31 U3)i + (T12Ul + T22U2 + T32 U3)i+


N

(T13U1

+ T23U2 + T33U3)k - q - qR -

L Pshsvs .

(30)

s=l
It may be noticed that the thermal effect of species diffusion is still present in the
global energy equation. However, when a local chemical equilibrium calculation
based upon curve-fit thermodynamic properties is performed, there is usually not
enough information available on the local composition to warrant an evaluation
of diffusion, which has to be neglected. This is not the case when the chemical
composition and flow simulation problems are coupled and solved simultaneously.

The closure to the mathematical problem of the governing equations for flows in
local chemical equilibrium will be provided again by thermal and caloric equations
of state, as shown in 2.2. With temperature and species composition known from
the equilibrium solver, the previous equations are tantamount to a general equation
of state relating pressure to density and internal energy

p = pep, e).

(31)

When curve-fit properties are employed, an expression such as Equation 31 is readily


available 5o ,51.

Computational Methods in Hypersonic Aerodynamics 171

4.3 Differential Form


Many practical applications of hypersonic flow theory are based upon the differential form of the governing equations. In regions where the flow is smooth, that
is no discontinuities such as shock waves or slip lines are present, application of
Gauss' Theorem to the integral form of the equations, Equation 22, written for a
fixed control volume, leads to
(32)
where the vectors Q, 8, 8 v , and W retain the form given in Equations 23 and 24
for thermochemical non-equilibrium, and Equations 28 and 29 for local chemical
equilibrium, provided that a Cartesian frame of reference is employed, and that the
boundary velocity Uv is set to zero.
When the flowfield over complex geometries is investigated, the limitations
of using fixed Cartesian grids become apparent, and a generalized nonorthogonal
reference system is usually needed 53. Moreover, unsteady problems involving moving boundaries lend themselves to simpler formulation and solution using timedependent, moving coordinate systems. Mathematically, the transformation between Cartesian and generalized frames is accomplished by writing the relations
~ =~(Xl' X2, X3,

t),
TJ =TJ(Xl, X2, X3, t),
(=((Xl, X2, X3, t),

(33)

where ~, TJ, ( represent a general structured coordinate system. For formulations


that employ unstructured meshes, the reader is referred elsewhere in this book.
Applying the chain rule for partial derivatives, after some algebra Equation 32 can
be written in generalized coordinates and in conservation form 54 as follows
(34)
where

W=J'
and

(35a,b)

F - Fv =,~~I [V't- (8 - 8 v) + ttQ]


-

G - Gv

IV'TJI[ _
=J
V'TJ (8 -

IV( I

8v)

_
+ 1JtQ] ,

H - Hv =j[V(. (8 - 8 v ) + (tQ].

(36a,b,c)

172 Computational Methods in Hypersonic Aerodynamics

In the equations above, J is the Jacobian of the coordinate transformation

(37)

'\It,

and
'\lit, '\l( are the unit vectors in the directions normal to the coordinate
surfaces

(38a,b,c)

Moreover, tt, itt, and (t are the normalized time derivatives of the transformation,
that is, the time derivatives divided by the magnitude of the gradient in their respective direction.
Denoting the components of the velocity in generalized coordinates, also known
as the contravariant components, as follows

(39)

inviscid and viscous flux vectors

F and Fv

become

PN U 1

-PNVN,

+ tXIP
PU1 U2 + {X2P
PU1 U3 + ~X3P

tX1711 +tx2 712 +tx3 713


tXI 721 +tX2 722 +tx3 723
tXI 731 +tX2 732 +tX3 733

PU1 U1

P1en , u1

PMe nM U1

pe OU1

+ pur

-qnl,1 -

PIe n, VI,

-qnM,1 - PMenMVM,

, (40a,b)

Computational Methods in Hypersonic Aerodynamics

173

where

In the above formulas, Vs stands for the species diffusion velocity written in generalized coordinates, whose components are obtained by a formula similar to Equation 39, but without the time derivative contributions. Similarly, the first component
of the generalized heat transfer vectors q, qR, and q~ are obtained by means of the
expresslOn

(42)
valid for the generic vector
of reference.

f,

whose components are known in the Cartesian frame

Similar results apply for the generalized inviscid and viscous flux vectors in
the other two space directions, when is replaced by r, and ( respectively, and the
proper components of it, V., q, qR, and in are selected. The expressions that are
valid for local chemical equilibrium can be obtained in an entirely similar way. The
final result for inviscid and viscous vectors F and Fv reads

, (43a, b)

where

e=

-iiI - iff -

L PShSVSl + (u1Fv2 + u2Fv3 + u3Fv4) I~~I .

(44)

s=l

The other two space directions are similar, once the proper components of velocity
and heat-flux vectors are selected.
A very interesting analogy can be drawn between the governing equations written in generalized coordinates and those written in integral form for an infinitesimal
control volume 55 . Inspection of Equations 34-38 reveals that the inverse of the J acobian of the transformation between Cartesian and generalized coordinates, 1/ J,
is geometrically equivalent to an infinitesimal volume. Moreover, the term /V k 1/ J,
where k stands for the generic generalized coordinate (~, 'T], or 0, is geometrically
equivalent to an infinitesimal area in the direction k. Consequently, the three space
partial derivatives can be interpreted as the first-order difference of fluxes in the ~, 'T],
and ( direction, respectively, which would account for the flux across the six faces of
an infinitesimal prism. Moreover, the normalized time derivatives of the generalized
coordinates, et, r,t, and (t respectively, are equivalent to components of the boundary velocity. Overall, the differential equations written in generalized coordinates

174 Computational Methods in Hypersonic Aerodynamics

are equivalent to integral equations written for an infinitesimal, prismatic volume.


This equivalence is exploited in conjunction with conservative finite-difference techniques, whereby the validity of the discretized differential equations is extended to
flowfields involving discontinuities due to their interchangeability with the "correct"
discretized integral equations.
The formulation examined involves moving boundaries. In the derivation of
these equations 56 , it is shown that a certain grouping of time and space derivatives
of Jacobian and metrics is zero analytically. This grouping goes under the name of
Geometric Conservation Law 57 , and reduces to a trivial statement for steady boundaries. It can be shown that enforcing the Geometric Conservation Law numerically
is very important for the accuracy of unsteady simulations 56 .
4.4 Basic Modeling of Viscous Fluxes
In order to complete the description of the governing equations for hypersonic
flows, additional quantities that appear in the formulation must be defined and/or
related to the conserved variables. In particular, the stress tensor, the heat-flux
vectors, and the diffusion velocities have to be modeled. In the following, some of
the basic results will be examined. For details on their derivation, the reader is
referred to Bird, Stewart and Lightfoot 58 , and Anderson 2 .
The most common assumptions that are made when dealing with transport
properties concern the treatment of the viscous stress tensor. Generally only N ewtonian fluids are considered, where there is a linear relationship between stress and
rate of deformation. Moreover, bulk viscosity is usually neglected, although it might
play an important role in cases involving rotational non-equilibrium 35 ,41. Under
these assumptions, the stress tensor components are expressed as

.. _
T'J -

J.L

(Guo.
GXj

GUO j
GXi

~ (GUO l
3 J.L Gxl

GUo2
GX2

GU03)8
GX3 'J

i,j=1,2,3,

(45)

where J.L is the viscosity coefficient and the symbol 8ij stands for the Kroneker delta.
For mixture of gases in thermal equilibrium, the heat flux vector q is modeled
by means of the product of a thermal conductivity coefficient k times the temperature gradient (Fourier's Law). The extension of this approach to flows in thermal
non-equilibrium is usually done by considering similar contributions from the nonequilibrium temperatures 21 ,37,39. The resulting expressions read

= -kVT,

s = 1, ... ,M.

(46a, b)

However, it should be pointed out that the validity of Fourier's Law for nonequilibrium conditions has not been fully assessed to date. The extension of this
model to Full Non-Equilibrium is usually done by introducing more contributions
from the electron/electronic temperature 21 ,28. More details are given by Lee l .

Computational Methods in Hypersonic Aerodynamics

175

The radiative heat-flux vector qR is modeled taking into account emission, absorption, and induced emission by means of a quasi-equilibrium approximation 2
A quasi-one-dimensional "slab" theory is used to simplify the analysis of the complicated integrals that ensue 59 . Modeling radiative heat transfer phenomena in
the context of thermo-chemical non-equilibrium is very much a subject of current
research 60 , especially because flows at Mach number in excess of 25 in general, and
atmospheric reentry conditions in particular, are such that convection and diffusion
are secondary forms of heat transfer when compared with radiation 2
The previous expressions involved quantities that are expressed in Cartesian
coordinates. When generalized coordinates are used, the space derivatives 8(-)/8x;,
for i = 1,2,3, need to be expanded according to the chain rule

(47a, b, c)

For many practical applications, one can make the fairly usual approximation of neglecting viscous contributions except in one space direction, normal to a
solid surface. The resulting equations are known as the Thin-Layer Navier-Stokes
equations 53 . The viscous flux vectors of Equations 24b and 2gb simplify because
only one component of each vector entry is nonzero. Moreover, when generalized
coordinates are used, the previous derivative expansions, Equations 47, reduce to
the following

where ~ has been considered to be the direction normal to the solid surface. Most
viscous computations performed in conjunction with thermal non-equilibrium are
actually Thin-Layer calculations, due to the excessive fine computational meshes
that would be necessary to actually resolve the cross-derivative terms that are neglected in the Thin-Layer approach.
More details on the modeling of viscous terms, and in particular on the treatment of diffusion velocities, will be discussed in 5.5.

176 Computational Methods in Hypersonic Aerodynamics


5. THERMAL NON-EQUILIBRIUM AND TRANSPORT PHENOMENA
In the previous section, the source vector entries associated with thermodynamic non-equilibrium were not defined. Moreover, the modeling of viscous fluxes
was left incomplete, including the problem of the determination of mixture viscosity
and thermal conductivity coefficients and species diffusion velocities in conjunction
with both equilibrium and non-equilibrium thermodynamics. In the present section,
some basic information is provided as to the current understanding and simulation
capabilities of thermodynamic non-equilibrium and its interaction with transport
phenomena. More details can be found in the literature that will be referenced.
5.1 Vibrational Non-Equilibrium
Two basic physical phenomena will contribute to the source terms Q. in Equation 23b, when it is assumed that vibrational non-equilibrium is present. The first
one is the change in vibrational energy caused by inelastic collision, that is creation or destruction of vibrating particles due to chemical reactions 35 , the second
one is due to elastic collisions, and is tantamount to energy exchanges between the
different internal energy modes. The most important contribution to the latter is
the relaxation of non-equilibrium vibrational energy towards the equilibrium levels at the local translational temperature l . This is usually modeled by means of a
Landau-Teller relaxation equation Z, with a characteristic relaxation time that is a
function of all the vibrating species present 2l , and that has been modified to account
for limitations of the original Landau-Teller theory 37. Additional contributions include energy exchanges among the different vibrating species, and exchanges with
the electron/electronic temperature l ,29. These terms are also modeled by means
of "standard" relaxation equations l , when deemed important. In many practical
applications, however, they are dropped and only one vibrational-electronic temperature is employed, because it is recognized that the previous energy exchange
mechanisms are very fast and efficient 28 .
5.2 Electron/Electronic Non-Equilibrium
In reentry aerodynamic applications 7 , the velocities involved are high enough
to warrant the presence of ionized species and free electrons behind the strong shock
waves. Moreover, the electronic modes of atoms and molecules can be out of thermodynamic equilibrium. In this instance it is usually assumed that a Boltzmann
distribution of electron/electronic energy levels is maintained at a common temperature T e , and the relaxation of this distribution to the local equilibrium values is
included in the modeling of thermodynamic non-equilibrium 2l ,24. Several assumptions and approximations are usually necessary to simplify the treatment of what
would otherwise be a flowfield with two different translational temperatures 33 . In
general, an electron/electronic energy equation is added to the system of equations
presented in Equation 22. Lee l , Park 37 , Appleton and Bray 6l, and others present
several possible versions of this equation, which differ mostly in the amount and variety of simplifying assumptions. It might be noted that special attention is needed

Computational Methods in Hypersonic Aerodynamics

177

when performing some of the simplifications, because in a few instances the governing equation might lose its invariance to an arbitrary Gallilean transformation 32 .
For cases where only one common vibrational/electronic temperature is employed, the treatment of electronic non-equilibrium reverts back essentially to what
has been briefly described in the previous subsection, and only one non-equilibrium
energy equation is present in the system given in Equation 22. The source terms
will then include non-elastic phenomena and vibration/electronic relaxation to the
equilibrium translational temperature. Kinetic energy and partial pressure of free
electrons are usually neglected 28 .
5.3 Rotational Non-Equilibrium
Many practical applications of hypersonic theory involve reentry flow problems.
Depending upon vehicle size and speed, there is an altitude regime in the atmosphere
where the Navier-Stokes equations cease to be accurate and the Burnett equations
become necessary for the simulation of thick shock waves 40 . In these cases, large
deviations from rotational equilibrium can occur 32 . Several investigators in recent
years have tried to model rotational non-equilibrium in the context of the continuum
hypothesis 41 ,62. Eppard and Grossman 41 present a simplified model for rotational
non-equilibrium, based upon the introduction of bulk viscosity, but show that overall
better results are obtained when the structure of shock waves is investigated by
means of a rotational non-equilibrium model. It may be useful to point out that
their non-equilibrium model can be represented as a special case of the Simplified
Non-Equilibrium Thermodynamic Model, and consequently the algorithms described
in the present study will be applicable to cases with rotational non-equilibrium.
5.4 Body Forces
Contributions due to body forces 2::s Ps9s appear in both momentum and energy equations, as shown in the vectors of source terms W of Equations 23b and 28b.
However, gravity is negligible for hypersonic applications, and the only physical
phenomena that are sometimes modeled are the forces due to both self-induced
and externally applied electric and magnetic fields 1 . These contributions become
important for ionized flowfield, as is the case in reentry applications. The electric
field is usually related to the gradient of the free electrons 37 . A few studies of the
interaction of magnetic fields with flowfields appear in the literature 63 . In general,
however, the consideration of electric and magnetic fields as part of the gasdynamic
source terms is complicated by the necessity of solving the Maxwell equations for
their determination. Again, this is subject of ongoing research efforts.
5.5 Transport Properties
The vector of viscous fluxes has been introduced in Equations 24b and 29b,
complemented by the expressions for the stress tensor components and the heatflux vectors, Equations 45 and 46. It makes use of several transport coefficients
that have yet to be determined, namely mixture viscosity coefficient fl, mixture

178 Computational Methods in Hypersonic Aerodynamics

thermal conductivity coefficient k, species non-equilibrium thermal coefficients kn"


and diffusion velocities Vs, which have to be defined and/or calculated from the
local flowfield properties.
The determination of mixture viscosity and thermal conductivity coefficients
can be broken into two steps: obtaining values for the single species coefficients
and using them to recover the mixture properties. Considering the species viscosity coefficient, the first step is usually accomplished by means of Sutherland's
formula 58 or Blottner's curve-fit values 64 , when simple, approximate methodologies
are deemed sufficient. Alternatively, collision integrals have to be evaluated starting from some accurate representation of the intermolecular potential, and used to
determine the species viscosity39. The thermal conductivity coefficient is usually
broken into contributions due to the different energy modes 21 . The modified Eucken approximation 39 is used to recover the final result as a weighted summation of
those contributions. The above procedure is followed for both thermodynamic equilibrium and non-equilibrium by using the relevant temperatures in the derivation.
Only the energy modes that are in equilibrium will contribute to the equilibrium
species thermal conductivity. The species non-equilibrium thermal conductivity kn,
is determined at this stage in a similar way, starting from the corresponding energy
contribution 21 . In the above, special attention has to be paid to ionized species and
free electrons 39 . The second step has been theoretically studied by Chapman and
Enskog for mixtures of monatomic gases 65 . However, the final result is too complicated and cumbersome for practical applications, and simplified semi-empirical
mixture rules are usually employed 47 to recover the mixture values J-L and k.
The species diffusion velocities Vs are complicated functions of several physical
phenomena, including gradients of concentration, temperature and pressure present
in the flowfield 58 . In general, simplifications are utilized to eliminate the minor
influences and obtain a manageable model based upon concentration gradients and
coefficients of binary and multi-species diffusion 2 . There are examples of simulations where less extensive simplifications have been performed in both reactive and
non-reactive flowfields 5 ,66, but at the present time the computational requirements
associated with reactive calculations are such that only simple models for transport
properties are truly practical.
5.6 Turbulence Modeling
Many flows of practical interest in the hypersonic regime are laminar, due to the
low densities associated with high altitudes 2 . However, transitional and turbulent
flows play an important role in the study and design of complete hypersonic aircraft
configurations. A discussion of the problem of turbulence modeling in conjunction
with reactive and non-equilibrium flows is beyond the scope of this Chapter. However, it may be useful to refer the reader to some efforts that have been initiated in
the field of algorithm development for turbulent flows 67 , whereby a unified approach
is developed for the solution of two-equation turbulence models fully coupled with

Computational Methods in Hypersonic Aerodynamics

179

the gasdynamic equations. The resulting algorithm has been obtained for a perfect
gas, however it is a direct extension of a Roe-type scheme of the kind that will be
discussed in the next section, and seems to be readily applicable to non-equilibrium,
reacting flows.
6. NUMERICAL FORMULATION
In order for a computer simulation to be possible, the governing equations presented in 4 must be discreiized, that is reduced to a set of algebraic equations to be
solved. A very convenient discretization approach is the finite-volume technique 55 ,
whereby the integral form of the governing equations is solved for the unknown volume averages of conserved variables in some small, but finite, control volume. The
advantage of this method is its use of the integral form of the equations, which allows
a correct treatment of discontinuities and also consistently treats general grid topologies. Other discretizations are also possible, like spectral techniques 68 , which have
been successfully applied to flow simulations in a chemically active environment 69 ,
and the more "standard" finite-difference approach 53 . Both of these alternative
discretizations are based upon the differential form of the equations, and special attention has to be paid to the proper treatment of discontinuities to avoid numerical
problems.
The governing equations in differential form have been written in conservation
form. Mathematically, this means that the coefficients of the derivative terms are
constant or, if variable, their derivatives do not appear in the equations. Moreover,
the divergence of some physical quantity can usually be identified in the equations.
This formulation generally leads to fewer numerical problems in the discretized version for flows with discontinuities. Related to the conservation form of the governing equations is the conservative property of a discrete algorithm 53 . Physically, this
means that the conservation principle that was used for the derivation of the equations is enforced exactly over volumes of arbitrary size within the domain of interest.
The finite-volume approach has the conservative property. Discretizing the integral
form of the governing equations with a conservative scheme yields the possibility of
capturing shocks and other discontinuities. Shock-fitting techniques have also been
attempted for flows out of chemical equilibrium 70-72, but their implementation is
at a relatively initial stage for general thermo-chemical non-equilibrium situations.
6.1 Finite-Volume Technique
The finite-volume technique takes advantage of the integral form of the governing equations to achieve a discretized algebraic approximation that can be handled
by a digital computer. The physical domain of interest is partitioned in a (large)
set of subdomains. In the following, the node-centered approach will be described,
whereby the volume and surface integrals that appear in the governing equations
are approximated in terms of the unknown volume-averaged values of each computational cell. This technique can be applied to arbitrarily shaped volumes, but

180 Computational Methods in Hypersonic Aerodynamics


will be illustrated for structured meshes, where a generalized curvilinear frame of
reference (~, T/, () can be defined and related through a continuous transformation
to a reference set of Cartesian orthogonal coordinates (Xl, X2, X3), as discussed in
4.3.
It should be noted that every infinitesimal contribution to the surface integrals
is either shared by two adjacent cells or belongs to the external boundary. Consequently, every approximation to the surface integrals which is independent of the
cell for which the contribution is considered will yield a scheme that has the conservative property, as discussed in the previous section. The reason for this is that
when an arbitrary control volume including several cells is considered, the contributions to the discretized equations which correspond to surface integrals that are now
inside the volume will cancel each other, leaving an equation that is the discretized
version of the conservation statement applied to this arbitrary control volume. In
the following, only discretizations with the conservative property will be examined.

The first step in the discretization process is to introduce volume-averaged


values for the conservative variables Q and the source terms W
(49a, b)
where the volume V of the generic computational cell is given by

and the Jacobian J of the coordinate transformation was defined in Equation 37.
Substituting into the governing equations in integral form, Equation 22, yields

8( <~; V) +

is

(S - Sv)' ndS =<W> V.

(51)

For the case of a structured mesh, the computational cell will be defined by two
= constant, two T/ = constant, and two ( = constant surfaces, respectively. Associating indices i, j, k to the ~, T/, ( coordinates, respectively, the arbitrary volume
under consideration will be described by the indices ij k and the surface integral will
be split into six contributions, one from each delimiting surface. Then area-averaged
values of a generic quantity f can be introduced
~

(52)
with similar definitions for the other indices, with the area Si+1/2 given by

(53)

Computational Methods in Hypersonic Aerodynamics


Specializing the previous results to f = (5 - 5 v ) n, with n being
the three space directions, respectively, the final result is obtained

a( <Q> V)

at

181

V[, Vij, V( for

+F - FVi+l/2Si+l/2 - F - FVi_l/2Si-l/2+
+G - GVj+l/ZSj+l/Z - G - GVj_l/ZSj-l/Z+
+H - HVk+l/2Sk+1/2 - H - HVk_l/2Sk-l/2

=<w> V,

(54)
where the orientation of the surfaces has been taken into account for the six contributions to the surface integral. The vectors F, Fv , have been defined in Equations 40
and 43, and their counterparts in the other directions are similar. In the above, the
components of the boundary velocity Uv in the ~-direction appears in the expression
for F, Equations 40a and 43a, as -[to It is important to point out that Equation 54
is still exact, because volume and area-averaged values have not been approximated.
However, at this stage it is necessary to relate the area-averaged values to the unknown volume-averaged values. The resulting set of algebraic equations will be the
discretized form of the original problem that is actually solved by means of a digital
computer.
An interesting analogy 55 between the finite-volume formulation discussed so far
and a more standard finite-difference technique can be obtained when the volume V
is written in terms of an average Jacobian and average increments in the coordinate
directions

(55)
Similarly, the areas can be written as follows

(56)
with corresponding results for the other terms.
Substituting Equations 55 and 56 into Equation 54, dividing by
dropping the average notation yields

aat (Q)
J

1 [ IV~I)

I ~~

(J

i+l/2(F - F v )i+l/2 -

IV~I

( J )i_l/2(F -

b.U~l.''1b.(

and

F v )i-l/2 +

1 [ IV1]I)
- IV1]1
- ]
+ b.1]
( J j+l/2(G - G V)j+l/2 - (J) j_l/2(G - G v )j-l/2 +
1 [ IV'(I
- IV'(I
- ] w
+ b.( (J) k+l/2(H - H v h+l/2 - (J) k-l/2(H - H v h-l/2 = J.
(57)
Equation 57 is the discretized form of the governing differential equations written
in generalized coordinates, Equation 34, that would have been obtained using a

182 Computational Methods in Hypersonic Aerodynamics

standard finite-difference approach. Consequently, the two methods are strictly


related, the main differences being the interpretation of the unknowns as volumeaverages and a consistent treatment of metric terms and Jacobians in the case of
the finite-volume strategy. This should not come as a surprise, since differential and
integral forms of the governing equations have been shown to be related for the case
of infinitesimal control volumes, as discussed in 4.3.
Relating the area-averaged values of inviscid and viscous fluxes to the volumeaveraged values of the conserved variable is the key operation of any finite-volume
strategy. A generalized extrapolation procedure to recover values at a face starting
from quantities at the center of a volume can be formulated for a generic vector q
as follows 55

(58)
with similar results for the other indices. In the previous formulas, I is the identity
operator, \7, t:. are backward and forward difference operators, respectively, and
q-, q+ are left and right extrapolations. The bars in the difference operators
indicate the presence of a limiter for the higher-order corrections. A review of
commonly used limiters and their properties is given by Sweby 73. The parameter
K, determines the extrapolation scheme used. A value K, = 1/3 yields a third-order
upwind-biased technique, a value K, = -1 produces the second-order upwind method,
whereas K, = 1 gives the classic second-order centered scheme.
Two different techniques have emerged in recent years for the treatment of the
inviscid flux vector. The first one identifies q with the flux vector itself. Consequently, a vector of inviscid fluxes is created from volume-averaged values of the
variables in a cell, and subsequently face values are created by extrapolation. The
second method, called MUSCL extrapolation, identifies q with the vector of conserved variables. The extrapolation produces face values for the conserved variables,
then the fluxes are constructed from the functional relationships

(59)
The latter technique is claimed to be more accurate than the former, at least for
perfect-gas flows 74 . It is useful to point out that for many applications the extrapolation of primitive or even characteristic variables is preferred to the extrapolation
of conserved variables, Q. Numerical experiments have shown that with this procedure more robust codes are obtained for high-speed flows in the case of both perfect
gas and reacting flows 8 . Specific techniques for the discretization of the inviscid flux
vectors will be discussed in 7.
The viscous vectors are usually discretized using standard second-order accurate
central differences for the second derivatives, whereas cross-derivative terms may be
treated in a diagonal-dominance-preserving fashion 75. Using central differences for
the inviscid vectors (K, = 1) yields unstable schemes, unless artificial dissipation is
added.

Computational Methods in Hypersonic Aerodynamics 183


6.2 Explicit Time Integration
After fluxes and source terms have been discretized and related to the vector of
unknown volume-averaged variables, the remaining problem to be solved is how to
advance the numerical solution in time, given a known initial state. Some algorithms
that were very popular even in recent years 53 use a coupled approach, whereby time
integration and space discretization are interconnected. However, the finite-volume
technique lends itself to a fully uncoupled procedure, the time integration being
segregated from the space discretization procedures.
Two classes of problems usually arise when dealing with time integration. The
first class contains the unsteady problems, where an accurate time integration is essential for the overall validity of the simulation. Moreover, some unsteady problems
could involve moving boundaries and/or moving and deforming control volumes 56 .
The second class involves steady problems, where time accuracy is of no concern.
A pseudo-transient problem is usually solved in this case, where the solution is
advanced in a pseudo-time until convergence to a steady-state is reached. Most algorithms will be able to handle both steady and unsteady problems, provided that
the discretized form of the Geometric Conservation Law is enforced in the calculation of the time derivative of the control volumes for cases when the control volumes
are deforming 57 .
A popular and computationally inexpensive scheme that is second-order accurate in time is the m-stage Jameson-style explicit Runge-Kutta method 8 . Other
explicit methods used in Computational Fluid Dynamics applications are discussed
in Fletcher 76. Unfortunately, all explicit schemes have the major drawback of becoming extremely inefficient for stiff problems, that is when some of the characteristic times associated with chemical reactions or thermal relaxation become much
faster than the fluid dynamics characteristic time 48 . In these conditions, the time
step necessary for stability can become even orders of magnitude smaller than the
value which would be necessary for an efficient resolution of the overall gas dynamic
transient 77. For this reason, time-implicit schemes are advocated, as they show an
increase in efficiency that can be dramatic 8 . Recently, hybrid schemes have been
investigated 69 , whereby the source terms are treated implicitly and the fluxes remain explicit. These techniques show some promise, but they are still the object of
current research.
6.3 Implicit Time Integration
Implicit time integration schemes are very popular especially for steady problems, due to their unconditional stability and the consequent increase in time step
per iteration and overall efficiency that they make possible 8 . Moreover, for reacting
flow problems it has been shown that treating the source terms in an implicit fashion
is tantamount to rescaling the governing equations so that all of the time scales in
a flow problem become of the same order of magnitude 77 . Consequently, implicit
algorithms are a viable solution to stiffness problems.

184 Computational Methods in Hypersonic Aerodynamics


For cases where time accuracy is important, as in unsteady problems, secondorder-accurate schemes in time have been developed, such as the implicit twostep Runge-Kutta scheme developed by Iannelli and Baker 78 , or three-time-Ievel
schemes 56 . Examples abound of problems where even first-order algorithms such as
Euler Implicit are deemed satisfactory 56. Moreover, for steady problems the Euler
Implicit technique is almost universally accepted as the most efficient approach.
After anyone of the implicit techniques is applied to the discretized equations,
a huge linear problem is obtained, whose unknowns are the volume-averaged values
at each control volume and whose left hand side comprises the Jacobian matrices of
fluxes and source terms. For three-dimensional problems, the storage requirements
for an exact solution of this linear problem are too restrictive. Approximate solutions
have been devised, namely solutions in a plane in conjunction with relaxation in the
third space direction 8 , or some sort of approximate factorization that reduces the
original problem to a series of smaller problems 56 . The effects of the availability of
vector and parallel architectures upon the choice of efficient algorithms is another
topic of current research.
The Jacobian contributions for the source terms can be obtained by analytical
evaluation of derivatives 27 . The J acobians for the inviscid fluxes are available as part
of the derivation of the algorithms 27 . In cases when flux-difference schemes of the
Roe-type are used, as will be discussed in 7, the Jacobians are complex to derive.
For perfect gas flows, approximate Jacobians have been developed 79 . An alternative
approach is to utilize the Jacobians produced by the flux-vector algorithms 80 , and
this is the most popular procedure. The viscous terms are treated in an approximate
fashion, using only the Thin-Layer contributions a.nd neglecting the cross-derivative
terms in the J acobians.
All of the techniques previously discussed apply to flows in local chemical equilibrium as well as flows in thermo-chemical non-equilibrium. The difference is the
size of the problem, due to the increased number of equations that are solved per
control volume in the latter case, and the fact that in the former case some thermodynamic properties have to be obtained from the composition solver or from
curve-fit expressions. Moreover, the evaluation of the inviscid Jacobian matrices
might become more complicated for local chemical equilibrium, due to the possible
presence of various "gammas" and their derivatives 34 . Assuming constant gammas
in the Jacobian derivation is common practice 20 , although not entirely justified by
the physics 2 ,34.
In order to reduce the computational requirements for thermo-chemical nonequilibrium flow problems, loosely coupled or even uncoupled techniques 48 have
been implemented, whereby the species continuity and the non-equilibrium energy
equations are solved separately from the gas dynamic equations (global continuity, momentum, global energy). It is not obvious that these algorithms yield an

Computational Methods in Hypersonic Aerodynamics 185


improvement in overall computational efficiency when compared with fully coupled approaches, because their robustness is diminished by the uncoupling process.
Again, the search for an optimum in efficiency and robustness is a subject of current
research.

7. FLUX-SPLIT ALGORITHMS
In recent years, flux-split techniques have obtained wide acceptance as an accurate means of discretizing the inviscid fluxes. Originally developed for a perfect
gas 9 - 13 , they have been extended to flows in chemical equilibrium 14 - 2o , and to
mixtures out of chemical and thermal equilibrium 21 - 27 . They are found to be very
accurate and robust when used for transonic, supersonic and hypersonic flows 74 , and
are fully compatible with conservative finite-volume, shock-capturing approaches.
Probably the two most popular schemes in the flux-vector splitting category
are the ones due to Steger and Warming l l and to Van Leer12. The basic idea is
to split the inviscid flux vector in one space dimension in two parts, containing the
information that propagates downstream and upstream, respectively. Then the two
parts are constructed using the extrapolation formulas of Equation 58 consistently
with the direction of propagation. When the MUSCL approach is used the result is

(60a,b,c)
where the downstream propagating flux P+ uses the left extrapolation Q- and
the upstream propagating flux P- the right extrapolation Q+. Alternatively, positive and negative fluxes are extrapolated directly using the negative and positive
extrapolation formulas, respectively.
Another very popular scheme, less robust for hypersonic flows 74 but more accurate, is the flux-difference splitting technique due to Roe 13 . It consists of an
approximate Riemann solver, where an arbitrary discontinuity is supposed to exist at the cell surface between the left and the right state, and an approximate
solution for this situation is written in terms of waves propagating upstream and
downstream. In this instance, the inviscid flux vector is not split, but reconstructed
from the upstream and downstream contributions, that constitute the left and the
right states in the Riemann problem.
Extensions to more space directions are usually made by superimposing pseudoone-dimensional problems, where the extrapolation formulas to get left and right
states keep track of the relevant direction only55. In recent times, a lot of effort
has been spent towards the design of "truly" multi-dimensional algorithms, which
would be at least approximately independent of the grid orientation 81 ,82. It is not
clear yet what the computational cost of the possible increase in accuracy would be,
and the topic is still subject of current research. It should be mentioned, however,

186 Computational Methods in Hypersonic Aerodynamics


that the preliminary algorithms published are for a perfect gas only, although the
extension to more complex thermophysical models is feasible. Moreover, alternative
formulation based upon grid adaptation 7o ,83 in conjunction with more "standard"
pseudo-one-dimensional techniques promise to achieve similar improvements over
fixed-grid calculations.

In the following, Steger-Warming-type, Van Leer-type, and Roe-type algorithms


for flows in thermo-chemical non-equilibrium and local chemical equilibrium will be
presented. The Simplified Non-Equilibrium Thermodynamic Model discussed in 2
will be at the foundation of the non-equilibrium techniques, due to the complications
that arise when two translational temperatures are introduced in a thermodynamic
model such as the Full Non-Equilibrium model. Attempts to derive algorithms for
multiple translational temperatures have been registered 33 , but are still preliminary
in nature 32 .

The algorithms to be presented do not exhaust the category of flux-split techniques. Osher-type formulations exist for perfect gases, and their extension to reacting flows has been attempted 84 . More recently, new kinds of flux splitting are
emerging85 , whereby the treatment of the pressure gradient plays a central role, and
they seem to show promise of improvement over the "older" schemes. Again, the
original development is for perfect gases.

7.1 Steger-Warming-Type Algorithm-Thermo-Chemical Non-Equilibrium


The flux vector splitting technique was originally developed by Steger and
Warming l l for perfect gases. It takes advantage of the homogeneity of the Euler equations to construct those portions of the inviscid fluxes that are associated
with the propagation of information in a given direction. The homogeneity property
carries over to flows in thermo-chemical non-equilibrium 27 .
The Jacobian matrices A = 8Fj8Q, B = 8Gj8Q, and C = 8iIj8Q, can be
diagonalized by means of right and left eigenvectors27. The diagonal matrix will
contain the characteristic speeds of propagation of information and, when split in a
non-negative and a non-positive contributions according to the sign of those characteristic velocities, will yield a splitting of the Jacobian in two parts, propagating
downstream and upstream respectively. Then the splitting of the inviscid flux into
a positive and a negative contribution is only a question of using the homogeneity
relation

F =

AQ =

F+

+ F- ,

(61a, b)

with similar results for the other space dimensions. The final result for the positive

Computational Methods in Hypersonic Aerodynamics

187

and negative contributions to the inviscid fluxes can be written as

PI

1'-1

=-_-AA

PN
PUI
PU2
PU3
Ple n,
PMe nM
ph o -pa 2/(i'-l)

PI
PN

PI

A
p(u2+tx2 a)
B
+ 21' p(u3+txa a)
ple n,

PN
p(Ul-~Xla)
A
c p(u2-tx2 a)
+ 21' p(u3-txa a)
Ple n,

PMe nM
p(h o + u~a)

PMe nM
p(ho -u~a)

p(Ul+~Xla)

(62)

where the non-negative and non-positive eigenvalues of the Jacobian matrix are
given by the following

A _IV'~I Ul -a lUI -al


J

.
(63a, b, c)
It is interesting to point out that simpler forms of the algorithm, including thermodynamic equilibrium and perfect-gas versions, can be recovered by simplifying the
equations in a fashion consistent with that discussed in 2.
C-

7.2 Steger-Warming- Type Algorithm-Local Chemical Equilibrium


The derivation of a Steger-Warming-type algorithm for flows in local chemical equilibrium is complicated by the fact that the governing equations are no
longer homogeneous of degree one 15 . This means that a positive or negative flux
cannot be constructed by means of a simple matrix multiplication of the corresponding Jacobian contribution times the vector of conserved variables, as was done
in Equation 61 b for the general thermo-chemical non-equilibrium case. An alternative derivation based upon convective "streams" yields a one-parameter family
of splittings that is still based upon non-negative and non-positive propagation of
information 19, as follows

(64)

where the isentropic index f plays a central role in the splitting. The non-negative
and non-positive eigenvalues are given by Equations 63, where the speed of sound
is the equilibrium value, defined in Equation 9, and f3 is equal to

(f - a)
f3 = f(f-l) ,

(65)

188 Computational Methods in Hypersonic Aerodynamics

where the parameter C\' is arbitrary. The value C\' = 0 allows this scheme to reduce
nicely to the perfect-gas form, and to be formally similar to its non-equilibrium
counterpart when the ratio of frozen specific heats l' is substituted by the isentropic
index r.
Other formulations have appeared in the scientific literature. Grossman and
Walters20 utilize the ratio of enthalpy to internal energy "(* to split the flux, and
approximate the speed of sound by substituting the same ratio for the isentropic
index, yielding (a*? = "(*p/ p. The final algorithm is formally identical to the one
shown in Equation 64 when C\' = 0 and "(* is substituted for r everywhere. Liou,
Van Leer, and Shuen 16 split only a portion of the fluxes, and add to both positive
and negative contributions one half of the unsplit portion. Again, their splitting
is not based upon the "true" sound speed. It may be useful to point out that no
conclusive evidence has yet been presented as to the superiority of one approach to
any other.

7.3 Van Leer-Type Algorithm-Thermo-Chemical Non-Equilibrium


An alternative flux-vector splitting has been developed for perfect gases by
Van Leer12. His formulation has continuously differentiable flux contributions and
has been shown \0 result in smoother solutions near sonic points than the StegerWarming splitting74. The derivation of the algorithm is based upon the splitting of
the fluxes by means of low-order polynomials in the components of the normalized
velocity parameter M = it/a (often referred to as a Mach number, even if it is a
vector). Then the mass flux for the vector P, which is PUI = paM1 , may be split
as PUI = f~ + f;;" where
(66)
The remaining fluxes may be split to yield

p = IV'~I f
J

PN/p

+ [Xl (-Ul 2a)/i'


U2 + [X2( -Ul 2a)/i'
U3 + [X3( -Ul 2a)/i'
Ul

Plenl/p

(67)

Computational Methods in Hypersonic Aerodynamics

189

where

f e

- h0

t (-iiI _ 2a)

c.,t

m (-Ul

a )2 ,

(68)

"(

and m is an arbitrary parameter. The value m = 1/(1'+1) allows the present scheme
to reduce nicely to the "classic" perfect-gas version, but numerical experiments
do not show any significant effect on the value of m. In the supersonic range,
p = P, p'f' == 0, when Ml > 1 and Ml < -1, respectively. Different authors
have developed very similar schemes 22 - 24 . The other space dimensions are treated
in a similar fashion.
7.4 Van Leer-Type Algorithm-Local Chemical Equilibrium
The derivation of a Van leer-type algorithm for flows in local chemical equilibrium has been successfully performed by several authors 15 ,16,19,20. The final result
given by Liou, Van Leer, and Shuen16 reads

(69)

where I;:' and I!' are given by Equations 66 and 68 respectively. The results presented by other authors differ in the choice of the parameter m in Equation 68,
although no strong argument has been provided in favour of any specific value.
Grossman and Walters20 develop a similar splitting, but again the enthalpy to iTlternal energy ratio "(* has to be substituted for the isentropic index r in all of the
expressions given, including the definition of an "approximate" speed of sound.

7.5 Approximate Riemann Solver-Thermo-Chemical Non-Equilibrium


The essential features of flux-difference split algorithms involve the solution of
local Riemann problems arising from the consideration of discontinuous states at
cell interfaces on an initial data line 1o . The scheme developed for perfect gases by
Roe 13 falls into this category and has produced excellent results for both inviscid
and viscous flow simulations 7 4.
At a cell interface, for a given time, it is possible to define a left state, Uf,
and a right state, (- )r, which correspond to positive and negative extrapolations of
cell-volume values, respectively, following the logic outlined by Equation 58. Then
a jump operator may be defined
[(.)] = C)r - (-)f .

(70)

The key step in the construction of an approximate Riemann solver 17 , involves


determining appropriate averages of eigenvalues, ~i' right eigenvectors, Ei , and

190 Computational Methods in Hypersonic Aerodynamics


wave strengths,

ai,

such that
N+M+4

[Q]

N+M+4

aiEi,

[F] =

;=1

ai ~i Ei ,

(71a,b)

i=l

for cell interface states which are not necessarily close to each other, so that [Q]
is arbitrary. In the above, a Cartesian component of the flux vector has been
considered, along with the eigensystem associated with its Jacobian matrix. The
results obtained will be extended to generalized coordinates by standard means 55 .
For this case, it is possible to define the eigenvalues
i = 1, ... ,N+M+2,
i = N+M+3,

(72)

i=N+M+4,
and the eigenvectors

Pi

o
o
o
o

Pi U1
Pi U2
Pi U3

, i = 1, ... , N,

Ej+N =

, j = 1, ... ,M,

(u~ -~i)Pi
(73a, b, c, d, e)
0

ih
P2

E N +M+1=

0
0
1
0
0

0
0
0
1
0

EN+M+2 =

U2

U3

PN

U1
U2
U3

en!

, k

= 3,4,

Computational Methods in Hypersonic Aerodynamics 191


along with the corresponding wave strengths
[Pi]

Cl'i

Cl'j+N

Pi -

[P]
0,2 '

[P]

[pje nj ]
A

e nj

~,

i = 1, ... ,N,

(74a)

j = 1, ... ,M,

(74b)

O:N+M+l = p[U2],

(74c)

O:N+M+2 = p[U3] ,

(74d)

k = 3,4.

O:N+M+k = 20,2 ([P] po'[Ul]) ,

(74e)

The solution of the approximate Riemann problem involves determining algebraic


averages p, ii, Pi, en" ho, ,(J;i, a, such that Equations 71 are satisfied. In the above,
the definitions
A

Pi =

p'

(-;:)

e ni =

(P-;;:'i)
-P- ,

(75a,b)

have been used, and 'ljJi represents pressure derivatives taken with respect to the i-th
species density and keeping the other conservative variables constant, according to
the formula

(76)
It is noteworthy that these averages are not unique. As pointed out by Abgra1l 25 ,
the algebraic problem posed in Equations 71 has multiple solutions, and different
values have been published in the literature for some of the Roe-type averages 22 - 27
The major differences between the different approaches is in the evaluation of averages of the pressure derivatives, or equivalently of ,(J;i, whereas the same results
are obtained for mass fractions and non-equilibrium contributions, if present. No
numerical evidence has been presented of the superiority of one scheme over the
others.
Following the derivation of Grossman and Cinnella27, and using the notation
(77)
for the arithmetic average, the necessary averages are determined to be

p =<.JP> 2 - ~[.JPt

(78a)

= .jPrPl,

<u.JP> - (U)r..;p;. + (u)..;Pi


- <.JP> ..;p;. + ..;Pi
,
p;j P).JP>
Pi =
<.JP>
'
A

U -

(78b)

~~~-~~~-

= 1, ... ,N,

(78c)

192 Computational Methods in Hypersonic Aerodynamics

j = 1, ... ,M,

(7Sd)
(7Se)
(7Sf)
(7Sg)

where averaged values of temperature, T, mixture frozen specific heat at constant


volume, c~, species equilibrium contribution to the internal energy, ei, ratio of frozen
specific heats, :y, and mixture gas constant, R, are defined by the following

T = _<_T--,-v'P_p_>

(79a)

<yIp> '
N

(79b, c)
i = 1, ... , N,

(79d)

(7ge)
(79f)

In Equation 79c, the integral averages of the species frozen heats at constant volume
C~i have been introduced, which in general will have to be numerically determined
with any of the available quadrature formulas. For cases where the simple trapezoidal rule is employed, it is easy to recognize that C~i reduces to the arithmetic
average of right and left values. Moreover, it may be noticed that all of the above
definitions reduce to the "usual" Roe-averages for a perfect gas model.
Rewriting Equation 71 b by grouping the repeated eigenvalues, simplifying the
result and transforming it to generalized coordinates 55 yields

(SO)
The [F]A term arises from the first N +M +2 terms of the sum in Equation 71b,

Computational Methods in Hypersonic Aerodynamics 193


corresponding to the repeated eigenvalue Ai

~l' and may be written as

[pN I p]
[Ul] - txJUl]
[U2] - tX2 [Ul]
[U3] - tX3[Ul]
[PI en! I p]

e
lt~ - a2 /er-I)
nM

(81 )
where

and Ul, Ul are trivially defined starting from the averages of the Cartesian components Ul, U2, {h
"*

'"

""

Ul =~XI Ul + ~X2 U2
"
. . . * ""
Ul =U l + ~t.

'"

+ ~X3 U3 ,

(83a,b)

The [Fh and [F]c contributions arise from the last two terms in the summation
of Equation 71b, corresponding to the eigenvalues iiI +a and iiI -a, and are found
to be

PN

Ul

tXI a

U2 tX2a
U3 tX3a

en!

(84)

194 Computational Methods in Hypersonic Aerodynamics


The approximate Riemann solver is implemented by computing the cell face
fluxes as a summation over wave speeds 10

(85)

which can be written as follows, using the previous results


(86)
with the absolute value of the wave speeds substituted in Equations 81 and 84.
When the MUSCL approach is employed, right and left states are evaluated using
the high-order interpolation formulas of Equation 58, applied to primitive or characteristic variables. When the flux interpolation approach is utilized, right and left
states correspond to the cell-center values immediately to the right and left of the interface under consideration. Equation 86 is only first-order accurate in this instance,
and high-order terms may be added 75 to render the scheme second or third-order
accurate. These additional terms will involve eigenvalues, right eigenvectors and
wave strengths, which have been given in Equations 72-74.

7.6 Approximate Riemann Solver-Local Chemical Equilibrium


The derivation of an approximate Riemann Solver for flows in local chemical equilibrium follows the same footsteps as the ones previously seen for thermochemical non-equilibrium. Again, the averages that enter the definitions of the
flux jumps at the interface are not unique 18 , and the non-uniqueness is reduced to a
choice between different formulas for the averages of pressure derivatives. Moreover,
even the choice of the two key pressure derivatives to be used in the formulation
is different for different authors. Vinokur 18 employs partial derivatives of pressure
with respect to internal energy per unit volume, pe, and density, whereas Glaister 17
prefers internal energy per unit mass, e, and density. Grossman and Walters 20 , on
the other hand, avoid the explicit use of pressure derivatives by means of the introduction of two alternative quantities, isentropic index r and enthalpy to internal
energy ratio
Correspondingly, three different expressions for the equilibrium
speed of sound are possible

,*.

a2

(op)
op

+h (
pe

op )

ape

(87)

and the definition of the Roe-averaged speed of sound will be affected by the choice
made.
Following Vinokur 18 , and using Cartesian coordinates, the starting point for
the construction of the approximate Riemann solver is again the determination of

Computational Methods in Hypersonic Aerodynamics

195

appropriate averages to be used in Equations 71 to determine jumps in conservative


variables and interface fluxes. These jumps will involve eigenvalues
i = 1,2,3,
i = 4,
i = 5,

(88)

eigenvectors

and wave strengths


0,1 =

[p] -

[~] ,

(90a)

P[U2] ,
= P[U3] ,

(90b)
(90c)

0,2 =
0,3

o,k =

2~2

([P] pa[ud) ,

k = 4,5,

(90d)

where the partial derivatives of pressure with respect to internal energy per unit
volume, 8pj8pe == K, and with respect to density, 8pj8p == X, have been introduced.
The solution of the approximate Riemann problem involves determining algebraic averages p, it, ho, X, r;" a such that Equations 71 are satisfied. The first three
averages are formally unchanged from the previous development, with density being
averaged geometrically and velocity and total enthalpy being defined by means of
"standard" Roe-averages. Vinokur1B suggests relatively involved formulas for the
pressure derivative averages, and determines the average speed of sound as follows

a2 = X+ (ho -

'2

~O)r;,.

(91)

Rewriting Equation 71b by grouping the repeated eigenvalues, simplifying the result
and transforming it to generalized coordinates55 yields Equation 80 again, where the
[F]A term arises from the first three terms of the sum in Equation 71b, corresponding to the repeated eigenvalue Ai = -5: 1 , and may be written as

196 Computational Methods in Hypersonic Aerodynamics


where

(93)
and -5. 1 , -5.; have been defined in Equations 83. The [Ph and [P]c contributions
arise from the last two terms in the summation of Equation 71 b, corresponding to
the eigenvalues -5. 1 +a and -5. 1 -a, and are found to be
1

{Xl a
U2 (X2 a
U3 ~X3a
,
,*
ho Ul a

Ul

(94)

The approximate Riemann solver is implemented exactly as discussed in the


previous subsection. It may be noticed that when the present formulation is used in
conjunction with a chemical equilibrium solver, the pressure derivatives are readily
available
X

_(8P)
-

8p

a2 -

_
hh-l)
,

pe

8p )

K,

== ( 8pe

_
= 'Y - 1.

(95a, b)

When a curve-fit formulation is used, the pressure derivatives have to be obtained by


analytical or numerical derivation 86 , or have to be provided alongside of the other
relevant pieces of information 50. In general, there is no guarantee of continuity
and/or smooth behaviour of the result obtained by numerical derivation of the
pressure curves 86
8. CONCLUDING REMARKS
Extensive numerical results have been obtained using the flux-split algorithms
presented. The reader is referred to some of the literature cited8 ,20,22,27,33,86 for
numerous examples of flow simulations in one, two, and three dimensions, inviscid
and viscous, laminar and turbulent, steady and unsteady, obtained in recent years
by the present and other investigators.
The algorithms described in this study have proven to be robust and accurate.
They have been implemented with considerable success into several production-level
computer codes, such as CFL3DE and GASP, which are available through NASA
Langley. Many essential portions of the feasibility study and design of the NASP
aircraft are being conducted using the technology described. However, there is still
a lot to be understood in the challenging field of hypersonic gasdynamics, and there
are many possibilities open for improvement of physical modeling and algorithm
efficiency. A few of the topics of current research have been mentioned throughout
the text, and undoubtedly the near future will see advances in some, or most, of
those areas.

Computational Methods in Hypersonic Aerodynamics 197

ACKNOWLEDG EMENTS
This work was supported in part by the US Air Force, Eglin AFB, Florida, under
Contract Number F08635-89-C-0208, in part by the National Science Foundation,
which funds the Engineering Research Center for Computational Field Simulation,
and in part by NASA Langley Research Center, under Grant NAG-I-776.
REFERENCES
1. Lee, J .-H. Basic Governing Equations for the Flight Regimes of Aeroassisted
Orbital Transfer Vehicles, Progress in Astronautics and Aeronautics, Vo1.96,
pp. 3-53, 1985.
2. Anderson, J.D.Jr. Hypersonic and High Temperature Gas Dynamics, McGrawHill, New York, 1989.
3. Palaniswamy, S., Chakravarthy, S.R. and Ota, D.K. Finite Rate Chemistry for
USA-Series Codes: Formulation and Applications, AIAA Paper No. 89-0200,
1989.
4. Hoffman, J.J. Development of an Algorithm for the Three-Dimensional FullyCoupled Navier Stokes Equations with Finite Rate Chemistry, AIAA Paper
No. 89-0670, 1989.
5. Drummond, J.P., Rogers, R.C. and Hussaini, M.Y. A Detailed Numerical Model
of a Supersonic Reacting Mixing Layer, AIAA Paper No. 86-1427, 1986.
6. Molvik, G.A. and Merkle, C.L. A Set of Strongly Coupled, Upwind Algorithms
for Computing Flows in Chemical Nonequilibrium, AIAA Paper No. 89-0199,
1989.
7. Gnoffo, P.A. A Code Calibration Program in Support of the Aeroassist Flight
Experiment, AlA A Paper No. 89-1673, 1989.
8. Walters, R.W., Cinnella, P., Slack, D.C. and Halt, D. Characteristic Based
Algorithms for Flows in Thermo-Chemical Nonequilibrium, AIAA Paper No.
90-0393, 1990.
9. Harten, A., Lax, P.D. and Van Leer, B. On Upstream Differencing and GodunovType Schemes for Hyperbolic Conservation Laws, SIAM Review, Vo1.25, No.1,
pp. 35-61, 1983.
10. Roe, P.L. Characteristic-Based Schemes for the Euler Equations, Annual Review of Fluid Mechanics, VoLl8, pp. 337-365, 1986.
11. Steger, J.L. and Warming, R.F. Flux Vector Splitting of the Inviscid Gasdynamic Equations with Applications to Finite-Difference Methods, Journal of
Computational Physics, Vo1.40, pp. 263-293, 1981.
12. Van Leer, B. Flux-Vector Splitting for the Euler Equations. Lecture Notes in
Physics, VoLl70, pp. 507-512, Springer-Verlag, Berlin and New York, 1982.
13. Roe, P.L. Approximate Riemann Solvers, Parameter Vectors, and Difference
Schemes, Journal of Computational Physics, Vo1.43, pp. 357-372, 1981.
14. Colella, P. and Glaz, H.M. Efficient Solution Algorithms for the Riemann Problem for Real Gases, Journal of Computational Physics, Vo1.59, pp. 264-289,
1985.

198 Computational Methods in Hypersonic Aerodynamics


15. Vinokur, M. and Liu, Y. Equilibrium Gas Flow Computations II: An Analysis
of Numerical Formulations of Conservation Laws, AIAA Paper No. 88-0127,
1988.
16. Liou, M.-S., Van Leer, B. and Shuen, J.-S. Splitting of Inviscid Fluxes for Real
Gases, NASA TM 100856, 1988.
17. Glaister, P. An Approximate Linearised Riemann Solver for the Three-Dimensional Euler Equations for Real Gases Using Operator Splitting, Journal of
Computational Physics, Vol.77, pp. 361-383, 1988.
18. Vinokur, M. Flux Jacobian Matrices and Generalized Roe Average for an Equilibrium Real Gas, NASA CR 117512, 1988.
19. Vinokur, M. and Montagne, J.-L. Generalized Flux-Vector Splitting for an Equilibrium Real Gas, NASA CR 117513, 1988.
20. Grossman, B. and Walters, R.W. Flux-Split Algorithms for the Multi-Dimensional Euler Equations with Real Gases, Computers and Fluids, Vol.17, No.1,
pp. 99-112, 1989.
21. Candler, G.V. and MacCormack, R.W. The Computation of Hypersonic Ionized
Flows in Chemical and Thermal Nonequilibrium, AIAA Paper No. 88-0511,
1988.
22. Grossman, B. and Cinnella, P. The Development of Flux-Split Algorithms for
Flows with Non-Equilibrium Thermodynamics and Chemical Reactions, AIAA
Paper No. 88-3595-CP, 1988.
23. Shuen, J.-S., Liou, M.-S. and Van Leer, B. Inviscid Flux Splitting Algorithms for
Real Gases with Nonequilibrium Chemistry, Journal of Computational Physics,
Vol.90, pp. 371-395, 1990.
24. Liu, Y. and Vinokur, M. Upwind Algorithms for General Thermo-Chemical
Nonequilibrium Flows, AIAA Paper No. 89-0201, 1989.
25. Abgrall, R. Extension of Roe's approximate Riemann solver to equilibrium and
nonequilibrium flows, in Note on Numerical Fluid Mechanics, Vol.29, pp. 1-10,
Vieweg, Braunschweig, 1989.
26. Wada, Y., Kubota, H., Ogawa, S. and Ishiguro, T. A Generalized Roe's Approximate Riemann Solver for Chemically Reacting Flows, AIAA Paper No. 89-0202,
1989.
27. Grossman, B. and Cinnella, P. Flux-Split Algorithms for Flows with Nonequilibrium Chemistry and Vibrational Relaxation, Journal of Computational
Physics, Vol.88, No.1, pp. 131-168, 1990.
28. Gnoffo, P.A., Gupta, R.N. and Shinn, J.L. Conservation Equations and Physical
Models for Hypersonic Air Flows in Thermal and Chemical Nonequilibrium,
NASA TP 2867, 1989.
29. Candler, G. The Computation of Weakly Ionized Hypersonic Flows in ThermoChemical Nonequilibrium, Ph.D. Dissertation, Stanford University, 1988.
30. Jaffe, R.L. The Calculations of High Temperature Equilibrium and Nonequilibrium Specific Heat Data for N 2 , O 2 and NO, AlA A Paper No. 87-1633, 1987.
31. Park, C. a:J.d Yoon, S. A Fully-Coupled Implicit Method for Thermo-Chemical
Nonequilibrium Air at Sub-Orbital Flight Speeds, AIAA Paper No. 89-1974,

Computational Methods in Hypersonic Aerodynamics

199

1989.
32. Grossman, B., Cinnella, P. and Eppard, W.M. Algorithms for Non-Equilibrium
Hypersonic Flows, in Fourth International Symposium on Computational Fluid
Dynamics, A Collection of Technical Papers, VoLl, pp. 443-448, University of
California, Davis, 1991.
33. Cinnella, P. and Grossman, B. Upwind Techniques for Flows with Multiple
Translational Temperatures, AIAA Paper No. 90-1660, 1990.
34. Cinnella, P. and Cox, C.F. An Efficient "Black Box" Solver for the Equilibrium
Composition of Real Gases, AIAA Paper No. 91-3322, 1991.
35. Vincenti, W.G. and Kruger, C.H.Jr. Introduction to Physical Gas Dynamics,
Robert E. Krieger, Malabar, 1986.
36. Clarke, J.F. and McChesney, M. The Dynamics of Real Gases, Butterworth
& Co. Ltd., London and Boston, 1964 (second edition under the title The
Dynamics of Relaxing Gases, 1975).
37. Park, C. Nonequilibrium Hypersonic Aerothermodynamics, John Wiley & Sons,
New York, 1991.
38. McBride, B.J., Heimel, S., Ehlers, J.G. and Gordon, S. Thermodynamic Properties to 6000 K for 210 Substances Involving the First 18 Elements, NASA
SP-3001, 1963.
39. Gupta, RN., Yos, J.M., Thompson, RA. and Lee, K.P. A Review of Reaction
Rates and Thermodynamic and Transport Properties for an II-Species Air
Model for Chemical and Thermal Nonequilibrium Calculations to 30000 K,
NASA RP-1232, 1990.
40. Fiscko, K.A. and Chapman, D.R Hypersonic Shock Structure with Burnett
-Terms in the Viscous Stress and Heat Flux, AIAA Paper No. 88-2733, 1988.
41. Eppard, W.M. and Grossman, B. Calculation of Hypersonic Shock Structure
Using Flux-Split Algorithms, ICAM Report No. 91-07-04, Virginia Polytechnic
Institute and State University, Blacksburg, 1991.
42. Treanor, C.E. and Marrone, P.V. Effect of Dissociation on the Rate of Vibrational Relaxation, The Physics of Fluids, Vo1.5, No.9, pp. 1022-1026,1962.
43. Marrone, P.V. and Treanor, C.E. Chemical Relaxation with Preferential Dissociation from Excited Vibrational Levels, The Physics of Fluids, Vo1.6, No.9,
pp. 1215-1221, 1963.
44. Candler, G. Translation-Vibration-Dissociation Coupling in Nonequilibrium Hypersonic Flows, AIAA Paper No. 89-1739, 1989.
45. Sharma, S.P. and Schwenke, D.W. The Rate Parameters for Coupled RotationVibration-Dissociation Phenomenon in H 2 , AIAA Paper No. 89-1738, 1989.
46. Park, C. A Review of Reaction Rates in High Temperature Air, AIAA Paper
No. 89-1740, 1989.
47. Bruno, C. Real Gas Effects, Hypersonics (Eds. Bertin, J.J., Glowinski, Rand
Periaux, J.), VoLl, pp. 303-354, Birkhiiuser, Boston, 1989.
48. Oran, E.S. and Boris, J.P. Numerical Simulation of Reactive Flow, Elsevier,
Amsterdam, 1987.

200 Computational Methods in Hypersonic Aerodynamics


49. Cinnella, P. and Cox, C. Robust Algorithms for the Thermo-Chemical Properties of Real Gases, in Fourth International Symposium on Computational Fluid
Dynamics, A Collection of Technical Papers, VoLl, pp. 216~221, University of
California, Davis, 1991.
50. Liu, Y. and Vinokur, M. Equilibrium Gas Flow Computations. I. Accurate and
Efficient Calculations of Equilibrium Gas Properties, AIAA Paper No. 89-1736,
1989.
51. Srinivasan, S., Tannehill, J.C. and Weilmuenster, KJ. Simplified Curve Fits for
the Thermodynamic Properties of Equilibrium Air, NASA RP 1181, 1987.
52. Oldenborg, R, Chinitz, W., Friedman, M., Jaffe, R, Jachimowski, C., Rabinowitz, M., and Schott, G. Hypersonic Combustion Kinetics, NASA TM 1107,
1990.
53. Anderson, D.A., Tannehill, J.C. and Pletcher, RH., Computational Fluid Mechanics and Heat Transfer, Hemisphere, New York, 1984.
54. Vinokur, M. Conservation Equations of Gasdynamics in Curvilinear Coordinate
Systems, Journal of Computational Physics, VoLl4, pp. 105~125, 1974.
55. Walters, RW. and Thomas, J.L. Advances in Upwind Relaxation Methods,
in State-of-the-Art Surveys on Computational Mechanics (Ed. Noor, A.K),
ASME Publication, 1988.
56. Janus, J.M. Advanced 3~D CFD Algorithm for Turbomachinery, Ph.D. Dissertation, Mississippi State University, 1989.
57. Thomas, P.D. and Lombard, C.K Geometric Conservation Law and Its Application to Flow Computations on Moving Grids, AIAA Journal VoLl7, No.10,
pp. 1030~1037, 1979.
58. Bird, RB., Stewart, W.E. and Lightfoot, E.N. Transport Phenomena, John
Wiley & Sons, New York, 1960.
59. Candler, G. and Park, C. The Computation of Radiation from Nonequilibrium
Hypersonic Flows, AIAA Paper No. 88-2678, 1988.
60. Cinnella, P. and Elbert G.J. Two-Dimensional Radiative Heat Transfer Calculations for Flows in Thermo-Chemical Non-Equilibrium, AIAA Paper No. 920121, 1992.
61. Appleton, J.P. and Bray, K.N.C. The conservation equations for a non-equilibrium plasma, Journal of Fluid Mechanics, Vo1.20, No.4, pp. 659~672, 1964.
62. Lumpkin, F.E.III, Chapman, D.R and Park, C. A New Rotational Relaxation
Model For Use In Hypersonic Computational Fluid Mechanics, AIAA Paper
No. 89-1737, 1989.
63. Palmer, G. The Effects of Self-Generated and Applied Magnetic Fields on the
Computation of Flow Over a Mars Return Aerobrake, AIAA Paper No. 91-1462,
1991.
64. Blottner, F.G., Johnson, M. and Ellis, M. Chemically Reacting Viscous Flow
Program for Multi-Component Gas Mixtures, Sandia Laboratories, SC-RR-70754, 1971.
65. Chapman, S. and Cowling, T.G. The Mathematical Theory of Non-Uniform
Gases, Cambridge University Press, Cambridge, 1970.

Computational Methods in Hypersonic Aerodynamics 201

66. Baysal, 0., Engelund, W.C. and Tatum, K.E. Navier-Stokes Calculations of
Scramjet-Afterbody Flowfields, presented at Symposium on Advances and Applications in Computational Fluid Dynamics, Chicago, 1988.
67. Morrison, J. Flux Difference Split Scheme For Turbulent Transport Equations,
AIAA Paper No. 90-5251, 1990.
68. Drummond, J.P., Hussaini, M.J. and Zang, T.A., Spectral Methods for Modeling Supersonic Chemically Reacting Flowfields, AlA A Journal, Vo1.24, No.9,
pp. 1461-1467,1986.
69. Eklund, D.R., Drummond, J.P. and Hassan, H.A. The Efficient Calculation of
Chemically Reacting Flow, AIAA Paper No. 86-0563, 1986.
70. Aftosmis, M.J. and Baron, J.R. Adaptive Grid Embedding in Nonequilibrium
Hypersonic Flow, AIAA Paper No. 89-1652, 1989.
71. Botta, N., Pandolfi, M. and Germano, M. Nonequilibrium Reacting Hypersonic
Flow about Blunt Bodies: Numerical Predictions, AIAA Paper No. 88-0514,
1988.
72. Valorani, M., Onofri, M. and Favini, B. On the Solution of Nonequilibrium
Hypersonic Inviscid Steady Flows, AIAA Paper No. 89-0671, 1989.
73. Sweby, P.K. High Resolution Schemes Using Flux Limiters for Hyperbolic Conservation Laws, SIAM Journal of Numerical Analysis, Vo1.21, No.5, pp. 9951011, 1984.
74. Van Leer, B., Thomas, J.L., Roe, P.L. and Newsome, R.W. A Comparison of
Numerical Flux Formulas for the Euler and Navier-Stokes Equations, AIAA
Paper No. 87-ll04-CP, 1987.
75. Chakravarthy, S.R., Szema, K.-Y., Goldberg, U.C., Gorski, J ..L and Osher, S.
Application of a New Class of High Accuracy TVD Schemes to the NavierStokes Equations, AIAA Paper No. 85-0165, 1985.
76. Fletcher, C.A.J. Computational Techniques for Fluid Dynamics, VoLl, SpringerVerlag, Berlin and New York, 1988.
77. Bussing, T.R.A. and Murman, E.M. A Finite-Volume Method for the Calculation of Compressible Chemically Reacting Flows, AIAA Paper No. 85-0331,
1985.
78. Iannelli, G.S. and Baker, A.J. A Stiffly-Stable Implicit Runge-Kutta Algorithm
for CFD Applications, AIAA Paper No. 88-0416, 1988.
79. Barth, T.J. Analysis of Implicit Local Linearization Techniques for Upwind and
TVD Algorithms, AIAA Paper No. 87-0595, 1987.
80. Liou, M.-S. and Van Leer, B. Choice of Implicit and Explicit Operators for the
Upwind Difference Method, NASA TM 100857, 1988.
81. Hirsch, C. and Lacor, C. Upwind Algorithms Based on a Diagonalization of the
Multidimensional Euler Equations, AIAA Paper No. 89-1958, 1989.
82. Dadone A. and Grossman, B. A Rotated Upwind Scheme for the Euler Equations, AIAA Paper No. 91-0635, 1991.
83. Wilson, G.J. and MacCormack, R.W. Modeling Supersonic Combustion Using
a Fully-Implicit Numerical Method, AlA A Paper No. 90-2307.

202 Computational Methods in Hypersonic Aerodynamics

84. Abgrall, R. and Montagne, J.L. Generalization of Osher's Riemann Solver to


Mixture of Perfect Gas and Real Gas, La Recherche Aerospatiale, No.1989-4,
pp. 1-13, 1989.
85. Liou, M.-S. and Steffen, C.J.Jr. A New Flux Splitting Scheme, NASA TM 104404,
1991.
86. Grossman, B., Cinnella, P. and Garrett, J. A Survey of Upwind Methods for
Flows with Equilibrium and Non-Equilibrium Chemistry and Thermodynamics,
AIAA Paper No. 89-1653, 1989.

- - ----------

Chapter 6:
Efficient Multigrid Computation of Steady
Hypersonic Flows
B. Koren, P. W. Hemker
Center for Mathematics and Computer Science,
P.O. Box 4079, 1009 AB Amsterdam, The Netherlands
Note: This work was supported by the European Space Agency (ESA), through Avions Marcel Dassault - Breguet Aviation (AMD-BA).

ABSTRACT
In steady hypersonic flow computations, Newton iteration as a local relaxation procedure and nonlinear multigrid iteration as an acceleration procedure may both easily fail. In the present chapter,
some remedies are presented for overcoming these problems. The equations considered are the
steady, two-dimensional Navier-Stokes equations. The equations are discretized by an upwind finite
volume method.
Collective point Gauss-Seidel relaxation is applied as the standard smoothing technique. In
hypersonics this technique easily diverges. First, collective line Gauss-Seidel relaxation is applied as
an alternative smoothing technique. Though promising, it also fails in hypersonics. Next, collective
point Gauss-Seidel relaxation is reconsidered and improved; a divergence monitor is introduced and
in case of divergence a switch is made to a local explicit time stepping technique. Satisfactory singlegrid convergence results are shown for the computation of a hypersonic reentry flow around a blunt
forebody with canopy.
Unfortunately, with this improved smoothing technique, standard nonlinear multigrid iteration
still fails in hypersonics. The robustness improvements made therefore to the standard nonlinear multigrid method are a local damping of the restricted defect, a global upwind prolongation of the correction and a global upwind restriction of the defect. Satisfactory multigrid convergence results are
shown for the computation of a hypersonic launch and reentry flow around a blunt forebody with
canopy. For the test cases considered, it appears that the improved multigrid method performs
significantly better than a standard nonlinear multigrid method. For the test cases considered it
appears that the most significant improvement comes from the upwind prolongation, rather than from
the upwind restriction and the defect damping.

204

Computational Methods in Hypersonic Aerodynamics

1. INTRODUCTION

1.1. Governing equations


The flow equations considered are the steady, two-dimensional (20) Navier-Stokes equations

pu

a
ax

pu 2 +p
puv

+~

ay

pu (e +..)
p

Re

aX

Tn

pv
pvu
pv 2 +p
pv(e+..)
p

Tn

=0,

(I a)

with

au 2 av
ax 3 ay'
au av
Tn= ay +~,
4 av
2 au
.=----3 ay
3 ax'
4
3

T=----xx

(Ib)

II

and with the total energy satisfying, assuming a perfect gas,


I

e = _ _ .L.... + -(u 2 +v 2 ).
y-I p
2

( Ic)

For a detailed description of the various other quantities used, assumptions made and so on, we
refer to any standard textbook. Suffice it to say here that these are the full, steady, 20, compressible
Navier-Stokes equations with as main assumptions made: zero bulk viscosity and constant diffusion
coefficients. (So, the flow is assumed to be laminar and its diffusion coefficients are assumed to be
temperature-independent.) For 1/ Re =0, diffusion has vanished and the remaining equations are the
Euler equations.
So far, real gas effects are not taken into account. The specific heat ratio y of the di-atomic gas
considered is assumed to be constant and determined by fully excited translational and rotational
energies only. (Though it could easily be replaced by a function ranging from zero up to the full
equipartition value, the vibrational energy is assumed to be zero.)
1.2. Discretization method
For a description of the basic computational method which is taken as a point of departure, we refer
to Hemker and Spekreijse [1,2,3] and Koren [4,5,6]. Here we give a concise overview of the main
characteristics only. For both the Euler and Navier-Stokes equations, we have considered first- and
higher-order accurate discretizations. Since in solving all types of equations our multigrid solution
methods are applied to the first-order discretized equations only (higher-order discretized equations
are solved in an outer, single-grid defect correction iteration), here we likewise limit the description of
the discretization to the first-order accurate one only.

Computational Methods in Hypersonic Aerodynamics 205


The Navier-Stokes equations are discretized in their integral form. The discrete system of equations is obtained by dividing the integration region Q into quadrilateral finite volumes Q,.} and by
requiring that the conservation laws, Equation (Ia) in integral form, hold for each finite volume
separately. At each volume wall, this discretization requires the evaluation of the convective flux vector and, additionally for Navier-Stokes, the diffusive flux vector.
For the evaluation of the convective flux vector we use an upwind approach, which follows the
principle of Godunov [7]. For the solution of the resulting I D Riemann problem, we prefer the
approximate Riemann solver of Osher and Solomon [8] in the P-variant of Hemker and Spekreijse [I],
this for reasons of both accuracy and efficiency (see Hemker and Spekreijse [I] and Koren [6]). The
left and right states in the I D Riemann problem (which determine the accuracy of the convective
discretization) are simply taken equal to those in the corresponding adjacent volumes, leading to
first-order convection accuracy.
For the evaluation of the diffusive flux vector, we use a central finite volume technique (see Koren
[6]). This technique is second-order accurate, but given the first-order accurate upwind discretization
of the convective terms, the overall accuracy remains first-order.

2. BASIC MULTIGRID SOLUTION METHOD


For the solution of the nonlinear system of first-order accurate discretized Navier-Stokes equations, in
the basic method a standard nonlinear multigrid technique is applied, with collective symmetric point
Gauss-Seidel relaxation as the smoother. The solution process is started by nested iteration. In the
relaxation method exact Newton iteration is used for the collective update of the four state vector
components in each finite volume. Nested grids are applied such that each finite volume on a coarse
grid is the union of 2X2 volumes on the next finer grid. Let QI, ... ,Q'-I,Q,.Q'+I, ... ,Qr be a sequence
of such nested grids. with Q 1 the coarsest and Qt. the finest grid, and let Nlq,)=r, denote the nona possibly non-zero
linear system of first-order discretized Navier-Stokes equations on Q" with
right-hand side related to the multigrid iteration. Then a single nonlinear multigrid cycle and the
nested iteration, as applied in the basic solution method, are defined in the following way.

r,

2.1. Nonlinear multigrid iteration


- Apply on Q, npre pre-relaxation sweeps to Nlq,)=r,.
- Compute the defect d, = N,(q,)- and restrict it to Q, -I: d, 1 = Ii 1d" where Ii - 1 is a restriction
operator for right-hand sides.
- Compute on the next coarser grid Q'_I the right-hand side r'_1 = N'_I(q'_I)-d'_I' For the initial estimate of q, I, we use the latest obtained q, -I'
- Approximate the solution of N, -I (q'_I) = r, 1 by the application of n a nonlinear multigrid cycles.
Denote the approximation obtained as - I '
Correct the current solution by q, = q, +Ii 1 (q, 1 - q, I). where Ii 1 is a linear prolongation operator for solutions.
- Apply on Q, n post post-relaxation sweeps to N,(q,) = r,.

r,

'0

For / = 1, the coarse grid correction is skipped of course. For the restriction operator Ii -I and
the prolongation operator Ii - 1 we take
(r, -I ),.; =

(Ii -I r, ),.; = (r;h, -1.2;

+ (r; hr - 1.2; + (r; b. 2; - 1 + (r,h,. 2;'

-,
_ -;
-,
-,
(/'-lq'-lh,-1.2;-I-(/'-lq, -lh,-1.2;-(/,- Iq'-lhr.2;-I-(/,
-

(2a)

Iq,

Ihr.2;-(q'I),,j"

(2b)

If not mentioned otherwise, for n a' n pre and n pm' we use in the basic multigrid method at each
level I: n a = I and n pre = n P'b' = I; i.e. as nonlinear multigrid cycles we use V-cycles with a single preand post-relaxation sweep per level.

206 Computational Methods in Hypersonic Aerodynamics


2.2. Nested iteration
- Choose a (possibly crude) initial estimate q I.
- Improve q 1 by a single nonlinear multi grid cycle as just defined above.
- Prolongate the improved approximation q 1 to Q2, yielding an initial estimate for q2'
- Improve q2 by a single nonlinear multigrid cycle as defined above.
- Continue the previous process until an initial estimate for qL has been obtained by prolongation of
qL-I'

The prolongation operator for obtaining the first approximation on each next finer grid may be
the piecewise constant operator (2b) or - preferably - a more accurate operator (for instance a 'bilinear
operator).
2.3. Numerical results
To give a quick impression of the performance of our basic multigrid method, both outside and inside
the hypersonic flow regime, we consider the following Euler flows: (i) the NACA0012-airfoil at
Moo =0.63, a=2 (smooth subsonic flow) and Moo =0.85, a= I (non-smooth transonic flow), and
(ii) a blunt forebody with canopy at Moo =8.15. a=30 (non-smooth hypersonic flow).
2.3.1. Subsonic and transonic airfoil flow. As finest finest-grid for the NACA0012-airfoil we consider
the 128 X 32 O-type grid given in Figure 1. In all corresponding multigrid cases, as coarsest grid we
consider the corresponding 8 X 2 O-type grid. (Hence for the 128 X 32-grid we have L = 5.)
For the present two airfoil flows the multigrid convergence histories are given in Figures 2a and
2b.
In both graphs, the residual ratio along the vertical axis is the ratio
Lt = II(N1Jq2k liLt = II(Ndq Lk I. L = 3.4. 5 versus the number of cycles performed; (i) one multigrid
cycle being a V-cycle with npre=npost=l, 't;fl, and (ii) for Q 5 only, one single-grid cycle being the
equivalent number of finest-grid relaxation sweeps. I(NL(q1Jh I denotes the summation - over all
volumes at Q L - of the absolute values of the k-th component in the first-order Euler defects, with q2
denoting the solution at Q L after the n-th multi- or single-grid cycle. Considering the corresponding
single-grid convergence histories, for both non-hypersonic cases. the effectiveness of the multigrid
method appears to be good.

a. In full.

b. In detail.

Figure 1. 128 X 32-grid (Q 5 ) NACA0012-airfoil.

Computational Methods in Hypersonic Aerodynamics 207

-------------------------125., _____ ---

10

cycleS
a, At Moo =0.63, a=2.
Figure 2. Convergence histories NACAOO 12-airfoil at 1/ Re = 0
( ------ : single-grid, - - : multigrid).

2.3.2. Hypersonic blunt body flow. However, for the hypersonic blunt body flow the results appear to
be different. The forebody is composed out of two ellipse segments (Figure 3), given by

l l
~

0.06

_x_
[ 0.035

2+
1

-----L. 12 - I
0.015

x<O,

2+ [-----L. J2 ~ I
0.025
J

(3a)

and a parallel part, given by


y= ~0.Q]5}
y=0.025

(3b)

0";;;;x";;;;0.016.

As finest finest-grid, we consider here the 64 X 32 C-type grid given in Figure 4. As coarsest grid, the
corresponding 4 X 2-grid is applied. (Hence, for this 64 X 32-grid we also have L = 5.)

0,025 - . - . - - - . -

------'1

o
-0,0\5 - -

T --. -=-=._-+--:
i
I

-0.06

0.016

Figure 3. Blunt forebody with canopy.

208 Computational Methods in Hypersonic Aerodynamics


~~-----------------------.

I<l

~+--------.-------.------~
-0.25

-0.15

-O.OS

0.05

x
Figure 4. 64X32-grid (rl s) blunt forebody with canopy.

With this gridding, with the basic computational method described before and with an initial solution q I which is equal to the upstream far-field boundary conditions, not any flow solution could be
obtained. Already in the first relaxation sweep on the coarsest grid, the solution process broke down!
A new research topic was found: extension to hypersonics of the basic method's applicability.

3. IMPROVED MULTIGRID SOLUTION METHOD


It can be quickly understood that even single-grid, hypersonic blunt body flow computations may
easily break down. In the initial phase of a steady flow computation in which strong solution perturbations arise, a local iterate may be easily swept out of the convergence range of the local Newton
iteration and cause global divergence. (It is obvious that this may easily happen in the very first visit
to the stagnation domain, during the computation of a hypersonic blunt body flow which has been crudely - initialized to its hypersonic upstream flow conditions and in which a strong shock wave is
arising.) Starting with a poor initial solution, one may gain in robustness by introducing a continuation process preceding the nested iteration. In such a process, usually a single upstream boundary
condition, for instance M co' is increased from some low initial value to its correct high value, while
performing relaxation sweeps. Continuation processes like this require a tuning of both the initial
value and the increment. For hypersonic flow problems, proper tuning is difficult because of the fact
that in these flows the condition numbers of the local derivative matrices used may be quite large.
(The larger the condition numbers, the larger are the perturbations in the iterates induced by perturbations in the right-hand sides; right-hand side perturbations which may already be quite large by
themselves in hypersonic flow computations.) The ill-conditioning occurring in hypersonic flow computations
can
be
quickly
illustrated
for
the
4 X4
Eulerian
derivative
matrix
:(pu, pu 2 +p, puv, pu(e +p/p, where ~ =(a/au, a/av, a/ac,_a/az)T, the differenti~1 ope:ator applied
III our solutIOn method, and where c
yp/p and z_ln(pp Y). Considerlllg for SimpliCity v =0 and
p = I, it clearly appears from Figure 5 that the condition of v (pu, pu 2 +p, puv, pu (e +p / p becomes
worse for u/c-'>oo. Notice further that the condition becomes worse also for u/c-'>O. The latter indi-

Computational Methods in Hypersonic Aerodynamics 209

10

Figure 5. Condition of typical derivative matrix to be inverted, v =0, p = I, II Re =0.


cates that stagnation flows become harder to relax with increasing upstream Mach number. (Further
investigations for II Re=O have shown us moreover that the poor performance of the point relaxation
method is not restricted to inviscid stagnation flows, but also holds for the very low speed flows
occurring in viscous sublayers.)
3.1. Line relaxation
If aforementioned situations are really of local nature, line relaxation may be a robust remedy. In a
local, very low-subsonic flow region such as e.g. a viscous sublayer adjacent to the wall, relaxation
lines crossing that layer and running into the outer solution (Figure 6a) are affected to a smaller
extent by the low speeds than single volumes in that layer. For a strong hypersonic shock wave arising in an initially unperturbed flow field, a similar reasoning may hold for relaxation lines crossing
that shock wave (Figure 6b) and single volumes in or downstream of that shock wave.
In a viscous sublayer with high aspect ratio volumes (such as in Figure 6a), an additional advantage of the properly directed line relaxation is that it is well-adapted to the corresponding strong coupling in crossflow direction. In convection dominated flow regions on the other hand, a strong coupling exists in flow direction. Here, relaxation lines are to be preferred which are more or less aligned
with the flow. So, if well-aligned, this is an additional advantage of lines crossing shock waves.
With line Gauss-Seidel relaxation as an alternative smoother, the basic multigrid method may in
principle be kept unchanged.

rl
I

Ml
---+-

a. Viscous sublayer.

b. Hypersonic shock wave.

Figure 6. Relaxation lines running out of difficult flow regions.

210 Computational Methods in Hypersonic Aerodynamics


3.1.1. Relaxation matrix. For line relaxation applied in an Euler flow computation, two basic types of
flows can be distinguished: flows with either (i) subsonic or (ii) supersonic velocity components along
the relaxation line considered. For the subsonic case, the upwind discretization scheme (correctly)
picks up its information from both upstream and downstream direction, with as a result: a block-tridiagonal relaxation matrix. For the supersonic case the result is a block-bi-diagonal matrix. For
Navier-Stokes flow computations a block-tri-diagonal matrix is the result in any case, except in the
rare case of supersonic velocity components and zero gradients of u, v and ("2 along the relaxation line.
(Then a blot:k-bi-diagonal matrix results again.) In all cases the blocks are 4X4-matrices. In implementing the line relaxation, we did not put any special effort into an efficient solution of a block-bior -tri-diagonal system; a solver for a general band matrix is applied.
3.1.2. Numerical results for supersonic flat plate flow. As test case for studying the convergence of
the multigrid method with line Gauss-Seidel relaxation as the new smoother, we start by considering a
supersonic flat plate flow with an oblique shock wave impinging upon the plate (Euler) or upon the
boundary layer (Navier-Stokes). The specific test case to be considered stems from Hakkinen et al.
[9]. It is the experiment at Moo = 2, Re = 2.96 105. Since in this experiment, the flow is known to be
laminar but yet hard to compute (because of the shock induced separation), it is a benchmark problem for laminar, 2D, compressible N avier-Stokes codes. A finest grid considered is the 80 X 32-grid
shown in Figure 7. In all multigrid cases the coarsest grid considered is the corresponding 5 X 2-grid,
hence for the grid in Figure 7 we have L = 5. The grids have been optimized for convection by a
stretching in flow direction and, in particular, by alignment with the impinging shock wave. A grid
adaptation for diffusion has been realized by a stretching in crossflow direction. The initial solution is
taken uniformly constant again and equal to the pre-shock inflow.
The convergence results are presented by the residual ratio Lk =11(NL(q7J)"I/Lk =11(N,Jq~)hl
versus either the amount of computational work (expressed in some appropriate work unit), or the
(wall clock) time. In the residual ratio, N L denotes again the discrete operator on the finest grid
(either first-order Euler or first-order Navier-Stokes) and q2 the iterate after the n-th work unit, with
q7. the iterate obtained by the nested iteration. All these flat plate flow computations have been performed on a (two-pipe) Cyber 205.
For the Euler flow, the multigrid behaviors for Gauss-Seidel relaxation with successively points,
crosswise lines and streamwise lines, are given in Figure 8. The streamwise line relaxation is symmetric whereas the other two relaxations are asymmetric with natural downwind sweeps only. To
ensure a good comparison of the various convergence rates, we define a work unit to be equal to: a
single multigrid cycle with symmetric relaxation, and consequently: two multi grid cycles with
downwind relaxation only. Clearly visible in Figure 8 are the expected superior convergence rates
and the better grid independence of the streamwise line relaxation.

Figure 7. 80 X 32-grid ([25) flat plate.

Computational Methods in Hypersonic Aerodynamics 211


In Figure 9, still for the Euler equations, for three different finest grids, the muItigrid efficiency
with the streamwise line relaxation is compared with that with the point relaxation. The markers
correspond with those in Figures Sa and Sc. Though no special effort was put into an efficient implementation of the line relaxation, it appears that its efficiency is the same for 1]3 and better for 1]4 and
1]5' (The gain in efficiency on finer grids is of course a consequence of the better grid independence.)
Also for the Navier-Stokes situation (Re = 2.96 105), we consider the multigrid behaviors with
point relaxation, crosswise line relaxation and streamwise line relaxation, successively. Here, all relaxations are symmetric, because of the occurrence of the subsonic sublayer. Further, the finest grid considered here is 1]7 and a work unit is defined as one multigrid cycle with symmetric relaxation. The
convergence rates are given in Figure 10. Both for point Gauss-Seidel relaxation and streamwise line
Gauss-Seidel relaxation we have divergence at 1]7. (For the latter relaxation we even have divergence
at 1]6') In both relaxation cases the cause of divergence is the increasing ill-conditioning directly above
the plate with decreasing mesh size normal to the plate. Here the crosswise line relaxation turns out to
be robust.

0'

ea';::l"'

.~

"0
'",

1]5
1]4

'l'

<1.l

--=OIl

1]5
1]4

1]3
1]2
1],

.2

10

1]]

........~I];

-~

+---~---''---~6---'--~lDl],

10

work units

work units

work units

a. Point relaxation.

b. Crosswise line relaxation.

c. Streamwise line relaxation.

Figure S. MuItigrid convergence histories for three types of Gauss-Seidel relaxation,


supersonic flat plate flow at M x =2, 1/ Re =0.

+---~------,---

2.5

7.5

time (sec)

12.5

JD

20

---

30

time (sec)

so

80

120

time (sec)
c. 1]5'

Figure 9. Multigrid efficiencies for three finest grids,


point (0) and streamwise line (e) Gauss-Seidel relaxation,
supersonic flat plate flow at M'X) = 2, 1/ Re = O.

J60

200

212 Computational Methods in Hypersonic Aerodynamics

--.--------.-----~-~

10

10

10

work units

work units

work units

a. Point relaxation.

b. Crosswise line relaxation.

c. Streamwise line relaxation.

Figure 10. Multigrid convergence histories for three types of Gauss-Seidel relaxation,
supersonic flat plate flow at M'l.) =2, Re=2.96 105.

An objection that can be made against the use of crosswise line relaxation throughout the computation is that though it is well-adapted to the strong coupling in the viscous sub layer with its high
aspect ratio cells, it is not well-adapted to the opposite coupling in the outer flow. (There, streamwise
lines are to be preferred.) The switch in direction of strong coupling suggests an adaptive local line
relaxation to be optimal (Figure 11).
In conclusion: for the supersonic Euler flow considered, streamwise line relaxation appears to be
most efficient. Already with a relatively slow solver for the large linear system, streamwise line relaxation may be more efficient than point relaxation. For Navier-Stokes flow computations with a practically relevant resolution of the viscous layers, crosswise line relaxation is to be preferred. Its advantage clearly is its greater robustness. It is less sensitive to a strong local ill-conditioning of the flow
equations.
3.1.3. Numerical results for hypersonic blunt body flow. Unfortunately, for the hypersonic blunt
body flow already considered in section 2.3.2, line Gauss-Seidel relaxation fails. In the next section we
return to point Gauss-Seidel relaxation and introduce a robustness improvement for it: a switch to
local, explicit time stepping.

-_._-----

--

----

~~j

t-~-l.'--~

x
Figure 11. Locally adapted relaxation lines.

Computational Methods in Hypersonic Aerodynamics 213


3.2. Switched-relaxation-evolution
In this section we do not yet consider the possibility of accelerating hypersonic flow computations by
muItigrid techniques. Here, we first restrict ourselves to the relaxation method only, particularly to its
robustness. Further, we also restrict ourselves to the Euler equations (1/ Re = 0).
As the remedy against failure of the Newton process in the point relaxation method, we use a
switched-relaxation-evolution technique. In this technique, we simply start applying the basic, collective point Gauss-Seidel relaxation method and take measures only as soon as the local Newton iteration fails. To discuss these measures for robustness improvement, we consider the local, first-order
discrete Euler system
I
<5;./q,)_T- (</>i H)F(T(</>i + 'c.,)q,./, T(</>, + ,,)qi + 1)1, + ".1I
T- (</>i - y,)F(T(</>i - ,.,.,)q, -I." T(</>, - '0./)q,)I, - ".j +
I
T- (</>,.j+ ,"JF(T(</>,., +,Jq,.J' T(</>,.,+ ,,)q,.j + I K, + ' i T-I(</>'.r !lJF(T(</>,.) - ,Jq,., -

J,

T(</>i.j - ,,)q,)I"I-" =0,

(4)

where T(</ denotes the matrix for rotation to a local coordinate system, F(ql, qr) the numerical flux
function (with its left and right cell face states q' and qr) and 1 the length of a finite volume wall. (For
further details we refer to Hemker and Spekreijse [1].)
3.2.1. Failing Newton iteration. As a non-failing Newton iteration to solve q,., from Equation (4) we
.
define: a Newton iteration for which: (i)

1("J,./q7./
I)h I
iii'
/I

1
",;;,

IUi)qi)kl

k = 1,2,3,4,

'!;I"J'

(5)

for any n-th Newton iterate (n =0, I, ... ,N) and each of the four residual components, and for which
I is physically correct, with physical correctness defined in the following way.
(ii) each iterate
Considering the local iterate q7./ =(u?j'
c?/, z;:)T and the corresponding hypersonic, upstream state
vector q 00 =(u Xl' v"" CXl' Z ",,)T, we know that the flow speed may not exceed the value corresponding
with adiabatic expansion to vacuum, starting from upstream conditions:

qU

<J'

'!;I,.}"

(6)

Further, we know that after this expansion, the speed of sound equals zero, its minimally allowable
value:
(7)

The maximally allowable value of the speed of sound is that corresponding with the stagnation temperature (which is the same for both isentropic and non-isentropic compression). For adiabatic flows
we can write:
/I C.
Ci'J~

2
ex

+ -2y- I (U 2oo + V 2

OO

'

(8)

For the lower limit of Z;'j we can directly write with the entropy condition:

z;:/ '?z x'

'!;I'f

(9)

For the upper limit of z?/ we have to consider the state q2 at the downstream side of a normal shock
wave which has at its upstream side a state q I which has expanded to vacuum, departing from
upstream conditions. Given the gas dynamics relations
P2=

2yMT -(y- I)
y+1
PI,

(y+ I)MT
P2= (y-I)MT+2PI,

(lOa)

(lOb)

214

Computational Methods in Hypersonic Aerodynamics


(lOc)

it is clear that
(11 )

Summarizing, we see that in adiabatic flows both the flow speed and the speed of sound have a physical lower and upper limit. The entropy only has a lower limit.
In the algorithm, the relations (5)-(9) are checked after each update in each local Newton iteration. As soon as one or more of these five requirements are not satisfied, that local Newton iteration
is said to have failed and the corresponding local correction found is rejected.
3.2.2. Evolution technique. As the alternative for a failing Newton iteration, we apply next one or
eventually two explicit time stepping schemes to the local, semi-discrete system
aqi.; +_I_Gi' (
)-0
A. Ji,j qi,i - .

at

( 12)

I';

with for qi,j here the conservative state vector (pl,j'(pu)"j'(PV)i.j,(pe)i,j)T and for A"j the area of finite
volume Qi-i'
As time stepping scheme to be applied first, we take the following version of the explicit, two-step
rational Runge-Kutta scheme of Wambecq [IOJ:
11+1_

qi.;

11

~Jfj(q7j)

T?j

"

I,'

-qi,j-w-:::;--2Gi'. (n )_Gi'. (" _1" ,II IA Gi'.. (


J',i qi,j
.J',i qi,j
;;2WT .j
i.j J , q,.j
"

n=O,I, ... ,N,

( 13)

<,

with W a possible damping factor for which we initially take w= I, and with
the local time step for
which we safely take the one which is maximally allowed for the forward Euler scheme:
A",h ,.j

T;: i = --[,.-':d<ff;)q;:j)
2.'--'-"''------0;-1 '

(14)

sup

dq',i

with h,) a characteristic local mesh size. With our upwind discretization, Equation (\4) may be rewritten by good approximation as
hi.;
T;:, = -,,
1(=u="=)=2=+::::::(V=I='=)2=-+-c-"V

/./

/./

( IS)

1./

For the evaluation of the denominator in scheme (\ 3) we use the Samelson inverse of a vector:

\1-1

1~1-2~,

(\6)

being a vector, whereas for the norm of a vector, we simply use the Cartesian inner product. As ini-

tial solution we take the same q7. i that just failed for the Newton iteration. The motivation for applying this scheme is its good stability as demonstrated in Wambecq [IOJ for a stiff and coupled system
of four equations, which is precisely what we have here in hypersonics. However, a potential danger
of scheme (\ 3) is that there is no guarantee for the denominator to be non-zero.
To protect Wambecq's scheme against a possibly too large time step and against a (nearly) zero
denominator, in each time step we require both the predictor and corrector to satisfy the conditions
(6)-(9). As soon as a physically unrealistic value occurs, the time stepping is stopped immediately,
rejecting any update made. Then, at first we assume that the unphysical result is due to a too large
time step. Therefore, as a remedy, we take w= lJ:! and restart the time stepping with Wambecq's
scheme, using the same q))1' In case of re-oCCurrence of something unphysical, we assume that the
denominator was the problem. Therefore, as a new remedy, we restart with an explicit time stepping
scheme which is safe in this sense; the simple forward Euler scheme

----

----------

Computational Methods in Hypersonic Aerodynamics 215


T"

q"1./+ I = q"1./ _ w---l.:.LuT


A. I.J.(q"1./.),

n -- 0 , I ,.,., N .

(17)

1./

For T7J in scheme (17) we also apply that according to relation (14). Further, for w we continue with
w =!/;? and for q?,j we also take the same as before. When a physically unrealistic value (according to
the conditions (6)-(9 occurs again, w is halved for the second time and the time stepping with forward Euler is restarted, still using the same q~j' In case of something unphysical once more, the time
stepping is stopped and the finite volume visited is left without any update being made. (Notice that
for both time stepping schemes, we do not require condition (5) to be satisfied.)
With the present switched-relaxation-evolution approach we expect that in those volumes where
Newton iteration fails, the local evolution technique will finally bring the solution into the attraction
domain of the Newton iteration (for the next sweep) and so make itself quickly superfluous.
3.2.3. Numerical results for hypersonic blunt body flow. To illustrate the benefits of the switchedrelaxation-evolution approach we consider again the hypersonic blunt body flow at M'l) = 8.15,
IX = 30.
In Figure 13, for the 16 X 8-, 32 X 16-, and 64 X 32-grids shown in Figure 12, we give the
corresponding convergence behaviors obtained by the switched-relaxation-evolution technique. The
residual ratio along the left vertical axes is again the ratio Lt = II(Nh(qh, I/LL II(Nh(q2, I. Here,
the solution q2 is the uniformly constant initial solution, which is equal to the hypersonic upstream
boundary conditions. (Notice that these are single-grid results.) The quantity along the right vertical
axes is the permillage of volumes in the total number of finite volumes visited during one cycle (one
cycle being defined as two diagonally opposite, symmetric switched-relaxation-evolution sweeps), In
which a switch to the evolution approach is made.
The robustness of the switched-relaxation-evolution technique is clear. For none of the cases considered is there an abortion of the solution process due to overflow or such. We even have convergence for all three cases. Further, from Figures 13b and 13c it appears that the evolution technique
makes itself superfluous indeed in the course of the iteration process. The extension to Navier-Stokes
can be quickly made. Only the time step needs to be reconsidered for diffusion. For this we refer to
e.g. Hindmarsh et al. [II].
~r-------------------

?+-----~-------r----~

-0."

-o.IS

a. 16 X 8-grid.

-D.~

b. 32X 16-grid.
Figure 12. Grids blunt forebody with canopy.

x
c. 64 X 32-grid.

0.05

216 Computational Methods in Hypersonic Aerodynamics

cycles

a. 16 X S-grid.

10

cycles

cycles

b. 32 X 16-grid.

c. 64 X 32-grid.

Figure 13. Single-grid convergence results switched-relaxation-evolution technique,


blunt forebody with canopy at Moo = S.15, IX = 30.

Notice that the convergence slow down with decreasing mesh size is expected for a plain relaxation method and is supposed to vanish by application of a suitable multigrid technique. Unfortunately, the switched-relaxation-evolution technique combined with the basic multigrid method as
described in section 2, does not lead to satisfactory results; see Figure 14, with here for q~ the
approximate solution obtained by the nested iteration. Standard changes to the multigrid algorithm,
such as for instance the replacement of V-cycles (n a = I) by W -cycles (n a = 2) do not help. It appears
that when applying multigrid to hypersonic test cases, in the standard way as described in section 2,
local coarse-to-fine grid corrections may be transferred which sweep the corresponding fine grid
iterates out of the attraction domain of the pure relaxation technique and even out of that of the
switched-relaxation-evolution technique. The cause of these problems may be either the coarse grid
corrections themselves, or the prolongation operator, or the combination of both. Therefore, in the
next section, to avoid possibly bad coarse grid corrections, we present a local damping technique for
the restricted defects and hence - implicitly - a local damping technique for the coarse grid corrections. To avoid a possibly bad correction transfer, in the next following section we also present an
alternative prolongation: a direction-dependent prolongation. The improvements have already been
published in Koren and Hemker [12]. For reasons of simplicity. here we keep ourselves restricted to
the Euler equations.

.,,
o

,+-----,----,r----,-----.----~

10

cycles
Figure 14. Convergence histories blunt fore body with canopy at Moo =S.15. 0:=0
( ------ : single-grid. - - : multigrid).

Computational Methods in Hypersonic Aerodynamics 217


3.3. Defect damping
3.3.1. Two-grid convergence analysis. De Zeeuw [13] reports a serious lack of robustness of standard
nonlinear multigrid applied to a test case described by the steady, I D semiconductor equations. De
Zeeuw meets the quest for greater robustness by a local damping of the restricted defect. Though the
way in which this damping is introduced by De Zeeuw is not very convincing from a theoretical point
of view (little evidence is given for the amount of damping to be applied). it certainly is convincing
from an experimental point of view. It leads to a significantly more robust multigrid method. For a
class of elliptic problems, a likewise successful, but basically different damping technique for improving the robustness of nonlinear multigrid, is that proposed by Reusken [14]. Instead of locally damping the restricted defect, Reusken proposes to damp globally (i.e. uniformly) the coarse grid correction, the amount of damping to be applied being prescribed by fairly rigorous theory. Yet, for the
present hypersonic flow computations, we prefer the type of damping as proposed by De Zeeuw, (i)
because it is not restricted to a specific class of elliptic problems (not even to elliptic problems in general), (ii) because of its a-priori character (a-posteriori damping like that of Reusken may already be
too late) and (iii) because of its local application (global damping like that of Reusken may strongly
reduce the positive effects of the coarse grid correction).
The local defect damping to be applied is now introduced by deriving the two-grid amplification
operator. Let
(18)
denote a nonlinear system of fine grid equations that we want to solve. Then the corresponding
+ I)-st coarse grid problem (n =0, 1, ... ,N) to be solved, reads

(n

( 19)

with S, ~ 1 denoting the operator for the defect damping, with ql' +', denoting the fine grid iterate as
obtained after the (fine grid) pre-relaxation, and with ql'~1 and ql'!i denoting the coarse grid iterates
before the (coarse grid) pre-relaxation and after the (coarse grid) post-relaxation, respectively. By
linearization (neglecting higher-order terms). from Equation (19) we derive the relation
11+1_ n

q'~1

q'~1

--

dN

(11 ) ~ 1
dN ( 11 + ")
'~I q'~1
S
I'~I
,q,
(11+"_ ')
d
'~I
d
q,
q"
q'~1
q,

01
N
n=, , ... , ,

(20)

in which qi denotes the fully converged solution of Equation (18). Considering for the coarse grid
correction q/' !II - ql' ~ 1 also the relation
j)~ 1 (q/,!/ -ql'~ d=(q/' +, -q/' + "),

n =0, I, ... ,N.

(21)

with ql' ++, denoting the fine grid iterate before the (fine grid) post-relaxation, it follows the two-grid
convergence result
n+ o' , _ ' - [1-1-'
q,
q,-,
'~I

ldN'~I(ql'~I)]--IS
j'_ldNMI'+")]( 11+"_ ')
d
'~I ,
d
q,
q"
q'~1

q,

n=O,I, ... ,N,

(22)

in which I, denotes the identity operator on Q,. From Equation (22) it is clear that in case the property
(23)

is satisfied, it holds
I ,, ~ 1

dN( n+!,)
dN( 11+")
, q,
( n + -" - ' ) - (I - S )1'- I , q,
(11 +
d
q,
q, - '~I
'~I'
d
q,

, '-

')

q, '

n=O,I, ... ,N,

(24)

which indicates that for optimal two-grid convergence no damping should be applied (S, ~ 1 = I, ~ I)'

218 Computational Methods in Hypersonic Aerodynamics


However, in hypersonic flow computations, qi _I and qi + ", may strongly differ from each other and
as a consequence also both lacobians in relation (23). For example, a hypersonic shock wave which
is detached on Qt, may easily be attached on Qt-I, with as a probable consequence that there, locally
relation (23) is not satisfied at all. If relation (23) is not satisfied, in particular if this is only very
locally the case, damping of the restricted defect at those places might be useful. For optimal twogrid convergence, from Equation (22) we derive as local damping factor for the defect in finite volume
(Qt-I),.J' to be applied in the (n + I)-st multigrid cycle:
n + I) - mm
.
(s t-I
'.J-

[I

II(N;-I )i-JII
[
] ,
max II(N;h -l.2j -III, II(N;hi -l.2tll, II(N;h,. 2j -III, II(N;h. 2j II

n=O,I, ... ,N,

(25)

with N;_I dNt-l(qi-d1dqt-l, with N; dNt<qi+I')ldqt and with 1111 some matrix norm. Notice
that the local damping factor (25) is more or less the 2D equivalent of the I D damping introduced by
De Zeeuw [13]. To see if some additional gain can be obtained by also allowing local defect
amplification, in a numerical experiment we will also consider
n+1

II(N;-di)1

(St-I)i.j-

[,

max II (Nth

,
-l.2j -III,

,],

n =0, 1.. .. ,N.

(26)

II (Nth -1.2)1, II(N t )2i. 2r III, II(Nth. 2jll

At convergence of the solution, the defect multiplication will also have converged, both in case of
multiplication (25) and in case of multiplication (26). However, as opposed to the correction damping
proposed by Reusken [14], the present defect multiplication will probably not have vanished at convergence, neither in case of (25), nor in case of (26).
3.3.2. Numerical results for hypersonic blunt body flow. We proceed with evaluating the two defect
multiplication techniques proposed. As the starting point for improvement we consider the multigrid
results given in Figure IS. The results have been obtained for the blunt forebody with canopy at
Moo =8.15, a=O, without any multigrid improvement and without nested iteration, but with the
switched-relaxation-evolution approach. The initial solutions q?, I = 1,2, ... , L are taken uniformly constant and equal to the hypersonic upstream boundary conditions. In this way we obtain a poorer initial approximation, but we have a more discriminating test problem and we are ensured of an unambiguous evaluation, since q?, I = 1,2, ... , L will be the same for the different multigrid improvements to
be considered.

--er
0

':;1

"a

='
:g

'-'

"',

.so ':'
~

10

cycles
Figure IS. Multigrid convergence histories blunt forebody with canopy at M ox; =8.15, a=O,
without nested iteration and without any multigrid improvement.

Computational Methods in Hypersonic Aerodynamics 219


With damping (25) as the defect multiplication, we obtain the convergence results given in Figure
16a. With the defect multiplication according to relation (26), both damping and amplification, we
obtain the convergence results given in Figure 16b. For the matrix norms, in both relation (25) and
relation (26), we applied the Frobenius-like norm

'(44

'2
]"
II(N,h.2J11=2 L.k>l.k,=I(N,h,.k,
'
(27)
2,.2)
the factor 2' simply accounting for the fact that in our case N,(q,) is a line integral form. Corresponding with the results in Figures 16a and 16b, in Figures 17a and 17b we show distributions of the multiplication operator SL -I,L = 5 (the multiplication factor distribution on [24), as applied in the last
(i.e. the IO-th) multigrid cycle. In both cases the local damping is confined to only the close neighborhood of the blunt body. Locally, in Figure 17b the damping appears to be a little bit stronger than
that in Figure 17a. However, globally this is more or less compensated by the local amplifications.
Notice that the maximal amplification factor that was found to be applied in Figure 17b is still 0(1)
only. The minimal damping factors in Figures 17a and 17b are much larger than those found by De
Zeeuw [13] for his specific nonlinear test case. A second difference is the good improvements found
by De Zeeuw [13] for his basic multi grid method's performance and the present modest improvements. Both differences suggest that (at least) for the present test case, in order to significantly
improve the results presented in Figure 15. defect multiplication is not needed as much as improved
grid transfer operators; the topic of the next section.

04

Os
,

Os

04

.-..

0'

.g

.~

~ ,

'il

"0

'@ "',
,go

::s

03

'-'

'-'

"',

03

,go
o

o
,+---~,----.----,-----.---~

10

,+-----,----.----,-----,----4
1

cycles

cycles

a. According to relation (25).

b. According to relation (26).

10

Figure 16. Multigrid convergence histories blunt forebody with canopy at Moo =8.15. a=O,
with defect multiplication.

220 Computational Methods in Hypersonic Aerodynamics


~

,;

N
C

,;

~~

..... 1l
0

0.9

\~,------1.

?+-------.------,------~~----~
-O.OB

-0.06

-0.04

-0.02

?+-------r------.--~--_r----~
-O.OB

0.00

-0.06

-0.04

-0.02

0.00

a. According to relation (25).

b. According to relation (26).

Figure 17. Distribution multiplication factors applied on Q 4 in the 10-th multigrid cycle.
blunt forebody with canopy at M % =S.15. 0:=0.

3.4. Direction-dependent grid transfer operators


3.4.1. Prolongation. The standard, piecewise constant correction prolongation may be illustrated as in
Figure IS. In mathematical terms. solution correction by means of the piecewise constant correction
prolongation may be written as

(q?ew l2, -

1.2) - 1

= (q)'ldb -1.2) - 1 + (tlq,_ 1)i.),

(q?ew h, - 1.2i

= (q)'ld hi - 1.2)

+(tlql-Il,.,.

(q?ewb. 2,- 1

= (q)'ld h,. 2) -

+(tlq'-I)i.1'

(q?ewl2,.2,

= (q)'ld b . 2i

(2Sa)

+(tlq'-I)i.,.

with
(2Sb)
The direction-dependent correction prolongation that we propose now can be illustrated as in Figure
19. In mathematical terms - analogous to formulae (2S) - solution correction by means of this
direction-dependent prolongation is written as

+!12 [(tlql-I)i-'/:,) +(tlq,_ I)i.,-,,],


+!12 [(tlq,_ I)i -"'.] +(tlql_ 1),-/ +'A],

(q?ewl2,. 2) (

q,new) 2,.2)

= (q)'ldh,. 2,_ ( Old)


- q,
2,.2)

+!12 [(tlq,_ I), + ",.) + (tlql_ 1),.)

'A,].

+!12 [(tlql_ I), +'1:.) +(tlql_ Ili.i

+,1

(29a)

with the four fine grid cell center corrections (Figure 19b) defined as central averages of the coarse
grid cell face corrections (Figure 19a). The coarse grid cell face corrections are defined by

Computational Methods in Hypersonic Aerodynamics 221


(~ql-I)i - 'h,) = (q7"-'1 ), - ';',} -(q71<i 1 ), - ",j'

( uq,
- I )i

)
+ 'h,j -- ( q,new)
- I i + 'h,j - ( qIold
- I ,+ ,",),

(~q'-I )i.j -

'h

= (q7,,-WI ),.j -

'h

(29b)

-(q71<i 1 ),,) -/;"

(~ql-I )"j + 'h = (q7,,-WI )i,) + 'h - (q71<i 1 )i,j + 'I"

where the cell face states are computed in an upwind manner. We notice that instead of the present
procedure of computing coarse grid cell face states in an upwind manner and from that, in a central
manner, fine grid cell center corrections, might as well have been the reverse: computing the coarse
grid cell face states in a central manner and from that the fine grid cell center corrections in an
upwind manner. However, a drawback of the latter approach is that it requires additional geometrical
data for the upwind computation of the fine grid cell center corrections.
I
I
(~/-l)i,j

I
I
I

(~/-l).,j

r-------l-------I
I

I
(~/-l)i,j

(~/-l)i,j

I
I

Figure 18. Coarse grid finite volume with corresponding next-finer grid volumes
and corresponding piecewise constant corrections.

(~/-l )i,j +IS

-;

-;

+
~

I
,-":
I

~
(~/-l)i,j-l;

Figure 19. Coarse grid finite volume with corresponding next-finer grid volumes
and corresponding direction-dependent corrections.

222 Computational Methods in Hypersonic Aerodynamics


Given a left and right cell face state (qleft and
state qface may be computed from

qright),

for a general ID upwind scheme, a cell face


(30)

f( q facJ = F( q left, q right),

where j(q) and F(qleft,qright) denote the exact and numerical Euler flux function, respectively. A
drawback of the Euler equations is that obtaining a primitive state vector like e.g. q=(p,U,V,p)T from
j (q)=(pu,pu 2 +p,puv,pu(e +p / p)l requires the solution of a quadratic algebraic equation. Fortunately, with the P-variant of Osher's scheme (see Hemker and Spekreijse [I]), for most Riemannproblem cases arising in aeronautics F(qleft,qright)= j(q.), q. being a well-defined, single state vector
on the wave path connecting qleft and qright in state space. Hence, with the P-variant, in most cases without evaluating F(qleft,qright) - we can directly identify qface as qface=q . For the O-variant of
Osher's scheme (see Hemker and Spekreijse [I]), in almost all Riemann-problem cases arising in
aeronautics F(qleft,qright) is found to be the sum of three different fluxes j(q). In these cases, because
of j(q)'s nonlinearity, the previous simple procedure is not possible. In the (rare) cases where the Pvariant also leads to a sum of fluxes, we solve the quadratic equation and in case of a positive
discriminant and one zero being physically irrelevant (negative p and/or p), we take the zero which is
physically relevant (positive p and p). In all other cases, we simply take qrace = 'l2(qleft + qright).
Because of the consistency of Osher's scheme at boundaries, there the present upwind prolongation
can also be applied in a consistent way. Notice that the upwind prolongation may lead to cell face
states which are local extrema in state space. In conclusion, we emphasize that by replacing the
piecewise constant prolongation operator by the present upwind prolongation operator, the complete
numerical method has become more consistent. Both the discrete Euler operator and the correction
prolongation operator are upwind now, both being based on the same upwind scheme: the P-variant
of Osher's scheme.
3.4.2. Restriction. A provable consequence of the upwind prolongation is that no restriction operator
I can be made for which the coarse grid finite volume discretization is a formal Galerkin approximation of the fine grid finite volume discretization. The possibly most effective restriction operator
that can be really made is the exact adjoint of the nonlinear prolongation operator. Unfortunatelyas opposed to the upwind prolongation - the exactly adjoint restriction operator will certainly lead to
a significant increase of the computational overhead. More suitable seems to be a linear approximation of the exact (nonlinear) adjoint. For this we write the latest obtained coarse grid cell face states
as linear combinations of the corresponding left and right states:

II

k = 1,2,3,4,

(qface)k =ak(qleft)k +(l-ak)(qright)k,

ak=

(qrace)k/(qright)k - I +12
(qleft)k/(qright)k - I

(3Ia)
(3Ib)

1,

where q is the conservative state vector, q=(p,pu,pv,pe)T, and a small parameter which guarantees
that (qface)k is a central average in case (qleft)k =(qrighth =(qraceh. With next the central computation
of the fine grid cell center states, we then have
(q/hi - I.2j -I.k =(i! -I q/-I hi - 1.2;-l.k

= 'l2(a/-I)i - Ye.j.k(q/-I)i -1.J.k +


'l2

[2 -(a/_I), - ",.j.k - (a/_I )i.j -'I,.k] (q/-I )i.j.k +

'l2(a/-I)i.j-~,.k(q/-di.)-I,ko

k = 1,2,3,4,

(32a)

(q/)z; -1,2j,k =(i! -I q/-I hr - I.2j,k = 'l2(a/-I)i -';,.),k(q/-I)i -1,J,k +


'l2 [I-(a/_ I ),- Ye,j,k +(a/_ d,,} +

'e,k] (q/-I)i,j.k +

'l2 [1-(a/_ d "jH,k](q/-I)i,}+I.k,

k =1,2,3,4,

(32b)

Computational Methods in Hypersonic Aerodynamics 223


(q,hi.21 -l.k

=(j~-Iq'-I h"

2j -

U:

= 0 [I-(a,_ di H"j,k] (q,-I)I+

l.j.k

0[1 +(a,_ di + !'.J,k -(a,_ d',r".k ](q,- di,j.k +


0(a'_1 )',j -'/,.k(q,-I )',j -l.b
(q,hi, 2j,k

=(j~ - I q,- d2"

2j,h =

(32c)

k = 1,2,3,4,

0 [I -(a'_I); H',j,k ] (q'-I); Hj,k +

o [(a,-I), +!,,/,k +(a'-I)i,j+'"k ](q'-I);'j,k +


o [1-(a,-d,,/+I",k](q'-I)i,)Hk'

k = 1,2,3,4,

(32d)

With for q the conservative state vector q =(p,pu,pv,pe l, the linear relations (32a)-(32d) display how
the upwind prolongation distributes mass, momentum and energy from a coarse grid to the overlying
finer grid, For the approximately adjoint restriction operator (i.e, the approximation of the exact nonlinear adjoint), we can then write
(r,_ d',j.k

=(I~-I r,),.j,k =

0 [1-(a"'1 ),_I,.J.k] [(r,h -2,2j -I,k + (r,)z, -2,2/.k] +

o [I -(a,_ d',j -'/"k]

[(rib - 1.2r2.k + (r')2;,

21

-2,k] +

0(a'_I), + '".),k [(r, b + 1.2/ -l.k + (r,hi + 1.2j,k] +


0(a'_1 );.1 + ",k

[(r,h,

-1,21

+ l.k + (rib. 21 + l.k ]

o [2-(a,_I), -!,.),k -(a'_I);.) - '!:.k ] (r,h, - 1.2) -l.k +


o [1-(a,_1 ),- ".j.k +(a'_1 ),.j +I".k] (r,b -1,2j,k +
o [I +(a,_I), + !',j,k -

(a'_1 )',j - I,.k ] (r,h" 2/ -I,k +

1;2 [(a,_ Il, H.j.k +(a'_1 )',j H'k] (r,b. 2j,k,

k = 1,2,3,4.

(33)

Of course, the weak spot in the approach (31 )-(32) is the linear approximation of the nonlinear prolongation. In case (qrace)k is a local extremum, i.e. does not lie in between (qleft)k and (qrighth, we
have a negative coefficient in relation (3Ia) (either ak or I-ak) and hence also in relation (33). We
do not accept this situation. If occurring, locally and for that k-th component only, we neglect how
the upwind prolongation really was and simply consider (qracelk = 0qleft)k + (qrightlk).
3.4.3. Numerical results for hypersonic blunt body flow, We now proceed with evaluating the
direction-dependent grid transfer techniques, As starting point for improvement we consider again the
multigrid results as obtained for the blunt forebody with canopy at Moo =8,15, a=O without any
multigrid improvement, but with the switched-relaxation-evolution approach, (See Figure 20, which is
the same as Figure 15.) The multigrid behavior obtained after having replaced both the basic correction prolongation operator and the basic defect restriction operator by the direction-dependent operators, is given in Figure 21a. (We remark that defect damping is not applied.) With upwind grid
transfers only, the improvement with respect to Figure 20 is significant indeed.
Replacing in the basic multi grid algorithm only the standard correction prolongation (by the
upwind prolongation), we obtain the multi grid performance given in Figure 21 b, These results are
only a little bit less good than those in Figure 21a and hence make us conclude that the previous,
rather cumbersome efforts in also upwinding the defect restriction, do not payoff enough. Therefore,
in the following we refrain from applying the upwind restriction,

224

Computational Methods in Hypersonic Aerodynamics

10

Figure 20. Multigrid convergence histories blunt forebody with canopy at M'l.) = 8.15, 0: =0,
without nested iteration and without any multigrid improvement.

,-.

.g

,-. ':'
0

.~

tIS

....
<a::s r

---

<a::s

D5
D4

~ 'f'

---

OIl

D3

OIl

D4

~ ,
~
..9

....

D5

..9

'l'

D3
'l'
c

cycles

10

a. With upwind correction prolongation


and upwind defect restriction.

b. With upwind

cycles

cor~ection

10

prolongation only.

Figure 21. Multigrid convergence histories blunt forebody with canopy at Moo =8.15,0:=0.

Computational Methods in Hypersonic Aerodynamics 225


3.5. Numerical results for hypersonic blunt body flow through combinations of multigrid improvements
3.5.1. Combination of defect damping and upwind correction prolongation. Though the two-grid convergence analysis in section 3.3.1 assumes that the prolongation operator is linear, see Equation (21),
no reason exists why local defect multiplication would have a detrimental effect in combination with
the nonlinear upwind prolongation. Therefore. in the present section, we show the multigrid performance for the combination of both defect damping and upwind correction prolongation. (Because
the results in Figure 16b already showed not to be better than those in Figure 16a and because of the
potential danger for divergence which is inherent to the allowance for local defect amplification, in the
following - for the defect multiplication - we apply relation (25). i.e. damping. only.)
Combining both multigrid improvements as considered separately in sections 3.3 and 3.4, we
obtain the results presented in Figure 22. Comparison with the results in Figure 21 b learns that the
combination of both techniques yields an only slightly better multigrid performance. We proceed with
further investigating the combination. In Figure 23a we show the distribution of the operator
SL-J,L=5, as applied in the last (i.e. again the lO-th) multigrid cycle; a cycle in which the solution
has already converged. First we notice that the damping has not vanished indeed. Further we notice
that though the solution must be symmetrical around the front ellipse, the damping factor distribution
is not. Cause of this is the fact that in the coarse grid problems, for the initial iterate, we take the
latest iterate computed. (See the description of the nonlinear multigrid iteration in section 3.3.1:
Equation (19).) By using the latest obtained iterate, the influence of the very first iterates is still felt,
iterates which - due to their poor level of convergence - are not yet symmetrical around the front
ellipse. An experimental proof of this explanation is given in Figure 23b in which we show the converged damping factor distribution for a strategy with solution restriction. Here, the converged damping factor distribution around the front ellipse is clearly symmetrical indeed. Notice that in both Figure 23a and Figure 23b the applied damping is modest. In Figure 23b - the case with solution restriction - it is even weaker than in Figure 23a. However. taking the restriction of the solution on the
coarser grids usually leads to a slower convergence than taking the latest available coarse grid iterates.

';"

'0'
'.::1

...
~

~
'0

'i...3

"-"

.,

Os
114

'",

.9 ,

03

'"
0

,
6

10

cycles
Figure 22. Multigrid convergence histories,
blunt forebody with canopy at Moo =8.15, a=O,
with both defect damping and upwind correction prolongation.

226 Computational Methods in Hypersonic Aerodynamics


c;

~
~
;..,~

Ei

o.jj)~~

,.

Ei

.;

...,~

(S4)min ROO. 15

~
~

Ei

p"'
{::
~ '(s.).;.~o."

~O'_
0.8

0.9

'0.7--

C;

C;

? -0.08

.;

~9

-0.06

-0.01

-0.02

a. Without solution restriction.

0.00

? -0.08

-0.06

-0.01

-0.02

0.00

b. With solution restriction.

Figure 23. Converged damping factor distributions on Q4,


blunt forebody with canopy at M Xi =8.15,0:=0,
with both defect damping and upwind correction prolongation.

3.5.2. Combination of defect damping, upwind correction prolongation and nested iteration. Comparing the basic multi grid method's results as presented in Figures 14 and 15 - results obtained with and
without nested iteration, respectively - clearly shows the natural beneficial influence of nested iteration.
For the improved multi grid method, the method with defect damping and upwind correction prolongation, the benefit of nested iteration (consistently, with upwind solution prolongation) is observed by
counting the number of finite volumes in which - locally - the switch is made from the relaxation
technique to the evolution technique (Figure 24). In both Figure 24a and Figure 24b (without and
with nested iteration, respectively), the quantity along the vertical axis is a scaled number of switches
made during the n-th multigrid cycle (n = 1,2, ... ,10), the scaling factor being the total number of
volumes visited during one nonlinear multigrid cycle; a V-cycle with npre=npost=1 and with symmetric relaxation sweeps. (Notice that the scaling factor increases when going from Q3 to Qd The
non-zero percentage at n = 0 in Figure 24b indicates the total amount of switches made during the
nested iteration. For all four grids considered, the expected positive influence of the nested iteration
appears to be significant.
In Figure 25 we give the multigrid convergence behavior corresponding with the latest favorite
strategy, the strategy with defect damping, upwind prolongation and nested iteration. In this figure,
for Q6, a comparison is also made with the corresponding single-grid convergence behavior. In Figure
26 we show the corresponding converged damping factor and Mach number distributions. Notice that
the smallest damping factors are mainly concentrated along the bow shock, in particular there where
the jumps across the shock are largest. Finally, we show results again for the more interesting reentry
case Moo = 8.15, 0: = 30. Also for this test case, the convergence results (Figure 27) show the
beneficial influence of the changes in the basic multigrid method. Given the very low convergence
rate of the single-grid computation (Figure 27a) and given the absolute failure of the basic multigrid
method (Figures 27a and 27b), the multigrid improvements do not just appear to be a nice luxury,
but a real necessity. Analogous to Figure 26, in Figure 28 we still show the converged damping factor
and Mach number distributions. Notice that - like in Figure 26a - the smallest damping factors in
Figure 28a are located at the most pronounced part of the bow shock.

Computational Methods in Hypersonic Aerodynamics 227

- - 0..

10

cycles

cycles

a. Without nested iteration.

b. With nested iteration.

10

Figure 24. Amount of volumes with switch from relaxation to evolution,


percentage of volumes visited during one multigrid cycle,
blunt forebody with canopy at Moo =8.15, a=O.

--- -- -- ------------ -- 0 6 - ---

.'0'
.,

':'
-......:-~--06

e ..,

:9

e'"

........ '

.2
':'
0

o.

10

cycles
Figure 25. Convergence histories blunt forebody with canopy at Moo =8.15, a=O,
with both defect damping, upwind prolongation and nested iteration,
( ------ : single-grid. - - : multigrid).

228 Computational Methods in Hypersonic Aerodynamics

~'---------------------~7r~---n

~.-------------------~

(!;

?+-------~------.-------.---~~
-0.08

-0.06

-0.01

-0.02

.oe

0.00

a. Damping factor distribution on

b. Mach number distribution on

Q s.

Q6'

Figure 26. Converged results blunt forebody with canopy at M YO =8.15, a=O,
with both defect damping, upwind prolongation and nested iteration.

\
"

"
"
"

,,
,, ,,

: ::
,,

f- :;

"

...

,+------.-------.------r-----~

10

15

cycles
a. Convergence histories.

20

8..

'

~',,--------- _____________________ _
10

15

20

cycles
b. Amount of volumes with
switch from relaxation to evolution,
percentage of volumes visited
during one multi-/single-grid cycle.

Figure 27. Iteration histories blunt forebody with canopy at M'l) =8.15, a=30,
with both defect damping, upwind prolongation and nested iteration, Q 6 only,
( ------ : single-grid, ........ : basic multigrid, - - : improved multi grid).

Computational Methods in Hypersonic Aerodynamics 229

;;

,;

N
0

,;

;..,~

;..,~

;;

;;

?+-------~------~--~~~~--~
-0.06

-0.06

-O.Oi

-0.02

0.00

?+-------.-------.-------.-0.06
-O.Oi
-0.02

a. Damping factor distribution on Os.

b. Mach number distribution on 0 6 ,

0.00

Figure 28. Converged results blunt forebody with canopy at Moo =8.15, a=30,
with both defect damping, upwind prolongation and nested iteration.

Concerning the efficiency of the improved multigrid method, one may find the paradoxical result
that one multigrid cycle with both defect damping and upwind correction prolongation is still cheaper
than one multigrid cycle without both. The cause of this simply is that the computations with the
improved multigrid method may result in a significantly smaller number of switches from the local
relaxation to a local evolution during the smoothing phases and hence in a lower computational cost.
Concerning the efficiency of the upwind computation of cell face states (as applied in the upwind prolongation), for the test cases considered it appears that with the P-variant, at almost all cell faces it
holds that F(qleft,qright)= f(q.). (For both 06-multigrid cases considered in this section, solving a
quadratic equation for q face appears to be necessary at about I % of all cell faces only.)

230

Computational Methods in Hypersonic Aerodynamics

To finish, we summarize the improved multigrid algorithm, the improvements being indicated in
bold.
Nested iteration:
- Choose ql.
- Improve q 1 by a single nonlinear multigrid cycle.
- Transfer the improved approximation q 1 to rl2' by applying the upwind prolongation operator.
- Improve q2 by a single nonlinear multi grid cycle.
- Continue the previous process until an initial estimate for qL has been obtained by upwind prolongation of qL ~I'
Nonlinear multigrid iteration:
Apply npre pre-relaxation sweeps to N/(q/)=r/.
Compute the defect d/=N/(q/)-r/ and restrict it: d/~I =I~~ld/.
Compute the local damping factors (S/ ~ 1 )',j and damp the restricted defect: d/ - 1 : = S/ ~ 1 d/ ~ I'
Compute the right-hand side r/ ~ 1 = N/ ~ 1(q/ ~ I) - d/ ~ 1 .
Approximate the solution of N/~I(q/~d=r/~I by the application of no nonlinear multigrid cycles.
Correct the current solution, by applying the upwind prolongation operator.
Apply npost post-relaxation sweeps to N/(q/)=r/.

4. CONCLUSIONS
For the hypersonic test cases considered in this paper, the essential element for robustness of the
smoother is the continuous monitoring of both the local relaxation and the local evolution. The
essential element for convergence appears to be the combination of Newton iteration, Wambecq's
explicit, two-step rational Runge-Kutta scheme and the explicit Euler scheme. For steady NavierStokes flow computations at a finite Reynolds number, the proposed checks on physical correctness
can be maintained as long as the flow remains adiabatic. Only the time step needs to be reconsidered
for diffusion.
A satisfactory remedy against divergence of nonlinear multigrid appears to be the combination of
a (local) damping of the restricted defect and a (global) upwind prolongation of the correction.
Besides a positive influence on the robustness of the algorithm, the combination of upwind prolongation and defect damping also has a positive influence on the computational efficiency. Applicationin addition - of an (approximately adjoint) upwind restriction operator does not really payoff. For
the test cases considered, the best improvement is obtained by the application of the upwind prolongation operator. With this operator we have achieved a greater upwind consistency throughout the
complete numerical method. For sake of clearness. we remark that the separate merits as observed
here for the multi grid improvements, may well be different for other test cases. Just as for the
switched-relaxation-evolution approach, the multigrid improvements are not restricted to the Euler
equations, but can be carried over as well to the Navier-Stokes equations.
Finally, we remark that the new techniques are such that the improved algorithm, just like the
basic algorithm, do not require any tuning of parameters.

Computational Methods in Hypersonic Aerodynamics 231


REFERENCES
1.

Hemker, P.W. and Spekreijse, S.P. Multiple Grid and Osher's Scheme for the Efficient Solution
of the Steady Euler Equations, Applied Numerical Mathematics, VoI.2, pp. 475-493, 1986.

2.

Hemker, P.W. Defect Correction and Higher Order Schemes for the Multi Grid Solution of the
Steady Euler Equations, in Lecture Notes in Mathematics (Ed. Hackbusch, W. and Trottenberg,
U.), VoI.1228, pp. 149-165, Proceedings of the Second European Conference on Multigrid
Methods, Cologne, 1985. Springer-Verlag, Berlin and New York, 1986.

3.

Spekreijse, S.P. Multigrid Solution of Monotone Second-Order Discretizations of Hyperbolic


Conservation Laws, Mathematics of Computation, VoL49, pp. 135-155, 1987.

4.

Koren, B. Defect Correction and Multigrid for an Efficient and Accurate Computation of Airfoil
Flows, Journal of Computational Physics, VoL77, pp. 183-206, 1988.

5.

Koren, B. Multigrid and Defect Correction for the Steady Navier-Stokes Equations, Journal of
Computational Physics, VoL87, pp. 25-46, 1990.

6.

Koren, B. Upwind Discretization of the Steady Navier-Stokes Equations, International Journal


for Numerical Methods in Fluids, Vol.Il, pp. 99-117,1990.

7.

Godunov, S.K. Finite Difference Method for Numerical Computation of Discontinuous Solutions of the Equations of Fluid Dynamics (Cornell Aeronautical Lab. TransL from the Russian),
Matematicheskii Sbornik, VoL47, pp. 271-306, 1959.

8.

Osher, S. and Solomon, F. Upwind Difference Schemes for Hyperbolic Systems of Conservation
Laws, Mathematics of Computation, VoL38, pp. 339-374, 1982.

9.

Hakkinen, R.J., Greber, I., Trilling, L. and Abarbanel, S.S. The Interaction of an Oblique Shock
Wave with a Laminar Boundary Layer, Memo 2-18-59 W, NASA, Washington, 1959.

10. Wambecq, A, Rational Runge-Kutta Methods for Solving Systems of Ordinary Differential
Equations, Computing, VoL20, pp. 333-342, 1978.
II. Hindmarsh, A,c., Gresho, P.M. and Griffiths, D.F. The Stability of Explicit Euler TimeIntegration for Certain Finite Difference Approximations of the Multi-Dimensional AdvectionDiffusion Equation, International Journal for Numerical Methods in Fluids, VolA, pp. 853-897,
1984.
12. Koren, B. and Hemker, P.W. Damped, Direction-Dependent Multigrid for Hypersonic Flow
Computations, Applied Numerical Mathematics, Vo1.7, pp. 309-328, 1991.
13. De Zeeuw, P.M. Nonlinear Multigrid Applied to a One-Dimensional Stationary Semiconductor
Model, SIAM Journal on Scientific and Statistical Computing (to appear).
14. Reusken, A.A. Convergence Analysis of Nonlinear Multigrid Methods, doctoral thesis, State
University of Utrecht, Utrecht, 1988.

Chapter 7:
Laminar-Turbulent Transition
D. Arnal
ONERA/CERT, Aerothermodynamics Department,
2 avenue E. Belin, 31055 Toulouse Cedex, France
ABSTRACT
The chapter is devoted to a description of theoretical, numerical and experimental problems related with
boundary layer transition at high speeds. On the theoretical point of view, emphasis is given on linear
stability theory which is able to take into account the effects of some parameters acting on transition. In
order to predict the transition location, the application of the so-called "en method" is discussed. Another
problem is to model the transition region. This can be done by using intermittency methods; some
comparisons with experimental results are presented. Other transition mechanisms are also described, in
two-dimensional (large roughness elements) as well as in three-dimensional flows (leading edge
contamination).
1 - INTRODUCTION
Since the classical experiments performed by Osborne Reynolds (1883), the instability of laminar flow and
the transition to turbulence have maintained a constant interest in fluid mechanics problems. This interest
results from the fact that transition controls important aerodynamic quantities such as drag or heat transfer.
For example, the heating rates generated by a turbulent boundary layer may be several times higher than
those for a laminar boundary layer, so that the prediction of transition is of great importance for hypersonic
reentry spacecraft, because the thickness of the thermal protection is strongly dependent upon the altitude
where transition occurs. Drag reduction for supersonic aircraft also requires a good knowledge of the
transition mechanisms.
When a laminar flow develops along a given body, it is strongly affected by various types of
disturbances generated by the model itself (roughness, vibrations ...) or existing in the freestream
(turbulence, noise, temperature fluctuations). These disturbances are the source of complex mechanisms
which ultimately lead to turbulence. Two kinds of transition processes are usually considered:
a) If the amplitude of the forced disturbances is small, one can observe at first exponentially growing
instabilities, the development of which is governed by ~ equations (Toll mien-Schlichting waves,
GlirtIer vortices, ... ). Three-dimensional and non linear effects occur subsequently, inducing secondary
instabilities and then transition. In these cases of so-called "natural transitions", the transition Reynolds
numbers can be very large.
b) If the amplitude of the forced disturbances is large (high freestream turbulence level, large roughness
elements, ...), non linear phenomena are immediately observed and transition occurs a short distance
downstream of the leading edge of the body. This mechanism is called a "bypass", in the sense that the
linear stages of the transition process are ignored (bypassed).

234 Computational Methods in Hypersonic Aerodynamics


The present paper is devoted to a general survey of transition phenomena in supersonic and hypersonic
flows, with emphasis on the numerical aspects of the problem. The subject is rather wide and several review
papers have described the state-of-the-art in the last years (Morkovin [I], [2], Reshotko [3], Mack [4], Pate
[5], Arnal [6], Stetson [7], ...). The first six sections deal essentially with two-dimensional problems.
Section 2 gives the main results of the linear stability theory, which can be applied in the case of "natural"
transitions. The next section describes the phenomena which occur upstream and downstream of the linear
phase; on one side, the receptivity theories explain how the forced disturbances excite the normal modes of
the laminar boundary layer; on the other side, non linear theories as well as direct numerical simulations
help to understand the ultimate mechanisms leading to transition. In section 4, it is shown that the linear
stability theory can be used to predict transition location in practical situations. The influence of various
factors is discussed in section 5 ; some of these factors modify the stability properties (wall cooling,
streamwise pressure gradient, bluntness, wall curvature) ; some others induce transition by a bypass
mechanism (large roughness elements). The problem of transition region modelling is addressed in section
6. The last section gives a short survey of peculiar phenomena related with three-dimensional configurations
(cross flow instability and leading edge contamination).
2 - FORMULATION OF LINEAR STABILITY THEORY
2.1. Historical backim>und
The instability leading to transition starts with the growth of sinusoidal disturbances, the existence of
which has been first demonstrated by the experiments of Schubauer and Skramstad (1948, [8]), in
incompressible flow. In fact, the development of small, regular oscillations travelling in the laminar
boundary layer was postulated by Rayleigh (1887) and Prandtl (1921) many decades ago. Some years later,
Tollmien worked out a complete theory of boundary layer instability (1929) and Schlichting calculated the
total amplification of the most unstable frequencies (1933). For this reason, the instability waves are often
referred to as the "Tollmien-Schlichting waves". A complete description of this pioneering work can be
found in Schlichting's book [9]. Nevertheless, the so-called "linear stability theory" received little
acceptance, essentially because of the lack of experimental results. The experiments performed by Schubauer
and Skramstad completely revised this opinion by demonstrating that T.S. waves constitute the frrst stage
of the transition process. In the last forty years, a large amount of work was devoted to a more and more
complete understanding of the linear mechanisms occurring at low speeds.
In compressible flow, the problem becomes more complex. The first attempt to develop a compressible
linear stability theory was made by Kiichemann (1938). One of the most important theoretical investigation
was carried out by Lees and Lin (1946, [10]), who used an asymptotic theory and deduced from the inviscid
equations some fundamental results related with the "generalized inflection point" (see below, paragraph
2.3.). Subsequent papers by Lees [11], Dunn and Lin [12], Lees and Reshotko [13] were also based on
asymptotic theories. However, as it was pointed out by Mack, "some major differences between the
incompressible and compressible theories were not uncovered until extensive calculations had been carried
out on the basis of a direct numerical solution of the differential equations." From 1960 to the present time,
Mack performed a thorough numerical investigation of the linear stability characteristics of compressible
laminar boundary layers [14] to [19] ; his most important finding is the discovery of the higher unstable
modes at supersonic speeds. Due to the development of high speed computers, extensive stability
computations have been performed during the last years. Some of the most significant results will be
presented in the next paragraph.
2.2. Stability eouations
A complete account of the linear stability theory is out of the scope of this paper (see Mack's review papers
[4], [15] for complete information). However, one needs to introduce some theoretical elements for a
comprehensive study of the experimental results.
The principle is to introduce sinusoidal small disturbances into the linearized state, continuity,
momentum and energy equations, in order to compute the range of unstable frequencies. It is assumed that
any fluctuation r' (velocity, pressure, density or temperature) is expressed by:
r' = r(y) exp [i (a.x +

~z

- rot)]

(1)

Computational Methods in Hypersonic Aerodynamics 235


The amplitude function r depends on y only: it is the so-called parallel flow approximation. a, ~ and
can be either real or complex quantities; in fact, two theories are essentially used :
in the temporal theory, a and

are real and

00

00

is complex, so that (1) can be written:

r' = r(y) exp(Olit) exp [i (ax +

~z

- Olrt)]

(2)

Oli represents the temporal amplification rate. Depending on its sign, the disturbances are amplified
(Oli > 0), neutral (Oli = 0) or damped (Oli < 0). Olr is a circular frequency. a and ~ are the components of
the wavenumber vector in the x and z directions, respectively. It will be shown that the wavenumber
direction 'If = tan-l(~/a) is a very important parameter.
in the spatial theory,

00

is real, a and

are complex. It follows from (1) that:

r' = r(y) exp [- (aix +

~iZ)]

exp [i (arx +

~rz

- Olt)]

(3)

We may defme a real wavenumber1 with an amplitude Ikl and a direction 'If :
Ikl=(a; +

~;r2

and 'If = tan-l

(~r/ar)

The spatial amplification rate is a vector with magnitude cr =

(a~

(4)

~ ~1 ) 1/2

and direction

'If = tan- 1 (- ~i/- ai) If ~i is set equal to zero, the waves can be amplified (ai < 0), neutral (ai = 0)
or damped (ai> 0).
As a general rule, Gaster's relation makes it possible to convert a temporal to a spatial amplification rate
by using the group velocity concept [20].
Introducing (1) into the linearized equations leads to a system of six ordinary differential equations in y
for the amplitude functions r(y). Due to the homogeneous boundary conditions (all fluctuations must vanish
at the wall and in the free stream, except the pressure fluctuations which have a non zero amplitude at the
wall), the problem is an eigenvalue one: when the mean flow is specified, a non zero solution exists only
for peculiar combinations of the four parameters a, ~, 00 and Re, where Re is the Reynolds number. In
fact, there are five real parameters in temporal theory (a, ~, Olr, Oli, Re) and six in spatial theory (ar , ai,
~

r, ~i,

00,

Re or k, 'If ' cr, 'If,

00,

Re).

From a numerical point of view, the input data are the boundary layer mean velocity and temperature
profiles, the free stream Mach number, the wall temperature, the characteristics of the fluid and all but two
parameters. The computation gives the two remaining parameters (eigenvalues) as well as the disturbances
amplitude profiles (eigenfunctions).

2.3. Inviscid theory


In this theory, it is assumed that the viscosity acts only on the mean profiles; furthermore, the terms of
the order lIRe are neglected in the stability equations. The mathematical study of the inviscid theory shows
up the importance of two parameters :
a) The first one is the so-called "generalized inflection point", which corresponds to the altitude Ys
where:

236 Computational Methods in Hypersonic Aerodynamics

II
dy

[dU]
p

= 0

(5)

dy ys

where p and u are the mean density and the mean velocity. It was demonstrated by Lees and Lin [10] that the
presence of such a point is a sufficient condition for the appearance of unstable disturbances; it is also a
necessary and sufficient condition for the existence of a neutral subsonic wave (the definition of a subsonic
disturbance will be given below) ; the phase velocity of this wave is the mean velocity at Ys. However,
these results are valid only if:
(6)

Let us consider the flat plate case. When the Mach number increases, the generalized inflection point
goes from the wall towards the outer edge of the boundary layer, and the numerical results indicate that the
range of unstable frequencies is enlarged at high Reynolds numbers. It is clear that this "inflectional
instability" plays a crucial role for instability and transition in hypersonics.
A

b) The second parameter is the "relative Mach number", M, which is defined as :

" = M-c/a

(7)

M is the local Mach number; for classical boundary layer profiles, it increases from 0 to Me between
the wall and the free stream. c is the phase velocity of the waves; it does not depend on y : for twodimensional waves (~ =0), c = wr/a in temporal theory, and c = w/a r in spatial theory. a represents the
local speed of sound, which depends on the mean temperature distribution; obviously, it takes a non zero

"

"

value at the wall. The disturbances are subsonic if M2 < I, sonic if M2

cia

= I, supersonic if M2
"

> I.

t
-1

"

Figure I-Typical evolution of the relative Mach number M


If a wave is locally supersonic, say between y = 0 and y = YM (figure I), the mathematical nature of
the stability equations changes, and any eigenvalue problem admits an infinite sequence of neutral
solutions; for instance, in the temporal theory and for two-dimensional disturbances, there is an infinity of
neutral waves (Wi =0), having the same phase velocity wr/a but different wavenumbers a. These
multiple solutions (the higher modes) were first discovered by Mack for boundary layer flows [14] ; in the
case of adiabatic wall conditions, they appear when the free stream Mach number exceeds 2.2.

2.4. Results for zero pressure gradient flow on adiabatic walls


The stability diagrams which are presented below were obtained by using a computer code developed at
ONERNCERT. The results are in good agreement with those published by other authors (at least for the
neutral curve). In these calculations, the basic profiles (mean velocity and temperature profiles) are obtained
by solving the two-dimensional compressible similarity equations. The real gas effects are not taken into
account, and the Prandtl number Pr is constant (Pr =0.725).

Computational Methods in Hypersonic Aerodynamics 237

0.4

ar

a,

0.3

t
0.2

0.1

_
0

510'

a) Me

ar

R61
0

10'

=0

R61

510'

b)

10'

= 1.3

Me

ar

_101ai

0.2

0.4
0.2

0.1

0.2

_
0

510'

a) Me

a,

R61

_
0

10'

= 2.2

510'

d) Me

ar

R61
10'

=3

0.5

0~______~===;~

__~--~~R:61~

10'

e) Me = 4.5

ar

ar

R61

0~-------------10~'------------~210'

210'

f) Me

= 4.8

r-_~10~'a-;~=ro-'--~=:==~=C~====~
4

<:;6

<6

__ R61

o~--~~~=-~~~~~
210'
410'

g) Me = 5.8

hJ Me = 7

ar

t6

, no =

_10

Figure 2- Stability diagrams computed for


two-dimensional waves (adiabatic wall)
_R61

00~-------'-------2-'-10'7"-----'----------.J104
4

i) Me = 10

238

Computational Methods in Hypersonic Aerodynamics

In the stability computations, the spatial theory is used, and it is assumed that the amplification vector
is aligned with the mean flow direction, i.e. ~i = 0 or 'II = 0 [21].
a) Two-dimensional waves. 'I' - 0
Figure 2 shows examples of stability diagrams computed for two-dimensional waves ('II = 0) and for
Mach numbers ranging from 0 to 10. For the sake of clarity, only some iso-amplification curves
(- ai > 0) are plotted in the (Reynolds number, wavenumber) plane. The Reynolds number is computed
from the displacement thickness BI, the free stream velocity Ile and the kinematic viscosity Ve of the outer
flow. a r and a i are made dimensionless with this displacement thickness. For each Mach number, there is
a critical Reynolds number (RBIcr) below which all disturbances are damped.
The first observation is that the stability diagrams do not change verJ much when the Mach number
increases from 0 to 1.3 (figures 2a-b). This can be explained by the fact that the generalized inflection point
remains very close to the wall in this Mach number range, and the boundary layer is unstable essentially
through the action of viscosity: this phenomenon, whereby the maximum amplification rate increases
with decreasing Reynolds number at a fixed value of ro or a r is called viscous instability; stability
diagrams where viscous instability is present exhibit a local maximum of - ai at fmite Reynolds numbers,
and the curves of equal amplification rate are closed around this peak. Another important observation is that
the amplification factors of the unstable waves are clearly smaller in the transonic range than in
incompressible flow, even if the critical Reynolds number remains of the same order.
At higher Mach numbers, the altitude of the generalized inflection point increases, and the inflectional
instability plays a more and more important role: the in viscid theory shows that there is a range of
unstable Clr or ro at infinite Reynolds number and this range is enlarged with increasing Mach numbers; as
a consequence, at large but finite Reynolds numbers, the curves of equal amplification rates tend to be
parallel with the RBI-axis.
For Me = 2.2 (figure 2c), the stability diagram clearly shows the viscous instability (below
RBI = 7 (00) and the inflectional instability (beyond RBI = 7 0(0). The latter becomes predominant for
Me = 3 (figure 2d).
It has been noticed previously that higher modes could be computed as soon as the free stream Mach
number becomes larger than 2.2, but they are at first associated with very large wavenumbers. For this
reason, only the first unstable mode is ploued in figures 2c and 2d for Me = 2.2 and 3. However, when the
Mach number is increased, the second unstable mode appears at lower and lower wavenumbers. This is
illustrated in figure 2e (Me = 4.5), where the first and the second modes are close together. Let us observe
that the second mode is more unstable than the first one: the ratio of the maximum amplification rates is
of the order of four of five. At Me = 4.8 (figure 2f), the two unstable regions join each other and there is
only one neutral curve in hypersonic conditions (figures 2g to 2i).
b) ObliQue waves. 'I' "# 0
The previous results were obtained for two-dimensional waves. However, an important aspect of
instability in compressible boundary layer is the effect of the wavenumber direction", on the amplification
rates. In incompressible flow, only two-dimensional waves need to be considered, because it can be
demonstrated that they are the most unstable ones (Squire's theorem). This is no longer the case in
compressible flow, even at moderate Mach numbers. In supersonic and hypersonic conditions, the situation
is particularly complex, as it can be seen in figure 3, where two stability diagrams are represented for
Me = 4.5 and two values of",: 0 and 60. Changing the wave orientation stabilizes the second mode,
but increases the instability of the first one, so that it becomes difficult to define "the most unstable
direction".

Computational Methods in Hypersonic Aerodynamics 239


w

t2

t2

0.5

lit' =0'1

oL-------~=====10~4==~--------~2~10~
_ R51
Figure 3- Effect of the wave orientation on the amplification rate (Me = 4.5)
c) Real gas effects
To date, most of the compressible stability analyses have assumed ideal gas behavior. Real gas effects
were studied quite recently by Malik [22], [23], Gasperas [24], Stuckert and Reed [25]. The latter study
demonstrates that the amplification rates associated with the second mode do not change very much from the
equilibrium to non equilibrium and ideal gas models, but that there is a shift in the range of the unstable
frequencies. The first mode is usually more stable than the second one, although it is strongly affected by
real gas effects.

3 - RECEPTIVITY, SECONDARY INSTABILITY, NON LINEAR EFFECTS


3.l. Receptivity
Some properties of the laminar boundary layer as a linear oscillator have been briefly discussed. But how are
the unstable waves excited by the available disturbance environment? This question is a part of the
problem which is usually addressed under the word of "receptjvity", introduced by Morkovin [1]. The
receptivity describes the means by which the forced disturbances (free stream turbulence, sound field, ... )
enter the laminar boundary layer; it describes also their signature in the disturbed flow. It has been shown
previously that a part of this signature is the development of unstable waves, which constitute the
eigenmodes of the boundary layer.
Recent works by Goldstein [26] and Kerschen [27] have demonstrated that the receptivity process occurs
in regions of the boundary layer where the mean flow exhibits rapid changes in the stream wise direction.
This happens near the body leading edge and/or in any region farther downstream where some local feature
forces the boundary layer to adjust on a short streamwise length scale (sudden change in the wall slope or in
the wall curvature, suction strip for instance).
These theoretical results were deduced from asymptotic approaches for incompressible flows; they have
been more or less substantiated by low speed experiments, but there is no general receptivity theory
applicable to high speed conditions. Mack [16] attempted to take into account the free stream noise effects
by developing a modified linear stability theory. Quite recently, Choudhari and Street [28] presented an
alternative to the asymptotic approach of Goldstein and Kerschen to predict the receptivity due to the
interaction of a free stream acoustic wave with localized regions of short-scale variations in surface boundary
conditions. This finite Reynolds number approach was applied to the generation of inviscid type
instabilities in a two-dimensional supersonic boundary layer.
It is now established that the receptivity increases (i.e. the initial amplitude of the unstable waves
increases for a fixed free stream Mach number and for a given disturbance environment) when do/dx
increases, where 0 is the boundary layer thickness. This could explain that the receptivity of a supersonic
boundary layer on a cone is less than that of a supersonic boundary layer on a flat plate. On the other side,
changing the unit Reynolds number in a given wind tunnel modifies the level and the spectrum of the free

240

Computational Methods in Hypersonic Aerodynamics

stream disturbances; this constitutes one of the possible reasons of the "unit Reynolds number effect",
which illustrates the importance of the receptivity process.

3.2. Non linear evolution and breakdown


Let us assume now that unstable waves were generated through the action of a certain receptivity process.
These waves grow according to the linear theory; when they reach a finite amplitude, the quadratic terms
neglected in the theory become appreciable, and non linear effects appear. The problem is fairly well
documented at low speeds, but very little is known about the non linear mechanisms in supersonic and
hypersonic flows. It can be guessed that the non-linearities are created by the temperature or density
fluctuations, the amplitude of which is several times larger than the velocity or pressure fluctuations.
Interesting information can be deduced from direct numerical simulations, see [29] to [31] for instance, as
well as from analytical theories [32].
The question arises of the length of the non linear zone at high speeds. In incompressible flows, the
distance between the end of the linear region and the breakdown to turbulence is rather short. This explains
that practical methods based on linear theory only (such as en method) provide us with good results for
predicting the transition location. Due to the lack of information, it is assumed that similar techniques can
be used in compressible flows, as it will be discussed in the following sections.

4 - TRANSITION PREDICTION
4.1. The en method
For a given melin flow, it is possible to compute a stability diagram (figure 4) showing the range of
unstable frequencies f as a function of the stream wise distance x. Let us consider now a wave which
propagates downstream with a fixed frequency f. This wave passes at first through the stable region ; it is
damped up to xO, then amplified up to x I, and it is damped again downstream of x I. At a given station x,
the total amplification rate of a spatially growing wave can be defined as :

x
In(NAQ) =

f-

(l

i dx

(8)

xO

~~------------~.x

Figure 4- Typical stability diagram in physical


coordinates. Definition of the total amplification rate and of the envelope curve

Computational Methods in Hypersonic Aerodynamics 241


A is the wave amplitude and the index 0 refers to the streamwise position where the waves become
unstable. The envelope of the total amplification curves is :

n = Max [In(NAQ)]
f

(9)

In incompressible flow, with a low disturbance environment, it is assumed that transition occurs as
soon as the n factor reaches a critical value in the range 7-10, Le. when the locally most unstable frequency
is amplified by a factor e 7 to elO [33, 34]. The problem is to know if the en method remains valid for
compressible boundary layers.
4.2. First application: flat plate with adiabatic wall
Envelope curves are plotted in figure 5 for four supersonic Mach numbers. We considered two-dimensional
waves (~r = 0), as well as oblique waves (~r 0) associated with the most unstable", angle (strictly
speaking, this angle depends on the Reynolds number for each Mach number, but its variation is rather
weak). Up to Me'" 5, the oblique waves are the most unstable ones; however, for hypersonic flows, the
two-dimensional waves become the most amplified again.

'*

Figure 6 shows an application of the en method for adiabatic flat plates. The stability diagrams were
used for computing the theoretical stream wise Reynolds numbers which correspond to different values of the
n factor, as a function of the freestream Mach number. The dotted lines represent theoretical results given by
Mack [16] for n =4.6 and 6, and the crosses are computations performed by Chen and Malik for Me = 3.5
and n = 2, 4, 6, 8 and 10 [35]. If it is assumed that transition occurs for a fixed value of the n factor, each
curve represents the evolution of the transition Reynolds number when Me increases. It can be observed that
compressibility has a strong stabilizing effect in transonic flow; it becomes destabilizing for Mach
numbers between 2 and 3.5, and stabilizing again in hypersonic conditions.
It has been noticed previously that the value of n at transition onset is usually between 7 and 10 in
"clean" subsonic wind tunnels. The problem is obviously to know if similar values of n are observed at
high speeds. The solid symbols in figure 6 represent experimental data obtained at ONERA some years ago
[36] : they correspond to low values of the n factor, between 2 and 4. Numerous examples could be given
for illustrating the fact that, in conventional hypersonic wind tunnels, the measured transition Reynolds
numbers are rather low, say between 2 and 3 millions. Possible origins of this behaviour are discussed
below.
4.3. Wind tunnel simulation - Empirical criteria
It is well known that transition on a smooth surface can be triggered by the disturbances which are present
in the free stream: velocity fluctuations u' (free stream turbulence), pressure fluctuations p' (acoustic
disturbances), temperature fluctuations T. The problem is very complex, because the effect of these various
disturbances depends on the Mach number range (Pate, [5]).
At low speed, the transition Reynolds number is very sensitive to the free stream turbulence level Tu ;
however, if Tu becomes very low, pressure fluctuations (fan noise for instance) can be of major importance
for inducing transition.
In transonic flow, acoustical phenomena linked with slotted or perforated walls give rise to strong
pressure disturbances, which cause early transitions. Better results are obtained in transonic wind tunnels
with solid walls.
In supersonic and hypersonic flows, the main factor affecting transition is also the noise, the origin of
which lies in the pressure disturbances radiated by the turbulent boundary layers developing along the wind
tunnel walls. The effects of u' and T' have not been firmly established for Me > 3.5.
The strong effect of the wind tunnel noise on the transition Reynolds number is illustrated in figure 7
(Harvey, [37]), where values of RXT, measured on cones and flat plates, are given as a function of

242 Computational Methods in Hypersonic Aerodynamics

n
10

/'
0

/'

t IMe =-2.21

IMe = 1.1 ]
5

n
10

/'

If! = 40 /
,//0 0
~

10

t IMe =

0
----00

10

-..10- 3 R~1

10

n
10

4.81

IMe = 5.81

--o~~----~------~---

20

20

40

Figure 5- Examples of envelope curves

10- 6 Rx

t
15

IhJ:.w.

n= 10 /

./

8//

Experiments

luillen
"quiet tunnel"

./

10

--

Mack
Chen and Malik
Present computations

./

5
Figure 6- Application of the en method
(flat plate, adiabatic Wall)

O~~~2~~--4~~--6~~-

-Me

Computational Methods in Hypersonic Aerodynamics

243

lrfl

RXT

8
6
4

,
Cones,

flight

107

8
6

\r
,

M,z5

lri'

8
6

(unflagged symbolsl

, "

Sharp flal plates


symhols L

(flag~ed

Sharp cones

Figure 7- Effect of wind tunnel noise on the


transition Reynolds number

li:

ata fairing

p~/qoo, for 4 < Moo < 23 (qoo is the mean dynamic pressure in the free stream and the - denotes a rootmean-square value). There are two separate mean curves for cones and flat plates, but the effect of Mach
number disappears completely in this representation: transition is governed by

p';' rather than by Moo.

Pate also analyzed available wind tunnels data and developed an empirical criterion for "natural"
transitions measured in supersonic and hypersonic facilities. This criterion is a correlation between the
transition Reynolds number RXT and the parameters acting on the noise intensity: the tunnel test section
circumference P and two characteristic parameters of the turbulent boundary layer on the nozzle walls (mean
skin friction coefficient CF and displacement thickness SO [5]. If the free stream Mach number and the wall
temperature ratio Twffaware constant, the criterion reduces to :
(10)

g increases as the test section circumference increases. This shows that the transition Reynolds number
increases with increasing unit Reynolds number UelVe and increasing wind tunnel size.
Another correlation, which was used during wind tunnel tests of a smooth model of COLUMBIA space
shuttle, is expressed by the simple relationship:
(R9/M e )T = constant, of the order of 200

(11)

R9 is the momentum thickness Reynolds number of the laminar boundary layer which develops on the
model. The values of R9T deduced from (11) are too low when compared with free flight data on smooth
models.
Since the radiated noise is inherent in the presence of walls around the model, there is little doubt that
wind tunnel conditions cannot simulate properly flight conditions. In order to reduce the noise level, it is
necessary to delay transition on the nozzle walls, because a laminar boundary layer is less noisy than a
turbulent one. This was done in the "quiet tunnel" built in NASA LANGLEY with a free stream Mach
number Moo = 3.5. A detailed description of the wind tunnel was given by Beckwith et al [38]. The nozzle
is two-dimensional and differs in design substantially from more conventional ones. Three notable features
are the use of a rapid expansion contour, a wide separation of the two planar side-walls and the provision of
boundary layer bleed slots shortly upstream of the nozzle throat. For unit Reynolds numbers between
2.5 105 and 15 105 per inch, the observed values of P~oo were found to vary from 0.03 10-2 up to about

244

Computational Methods in Hypersonic Aerodynamics

0.8 10-2 , depending upon the unit Reynolds number, the axial location and the bleed slot flow (bleed valve
open or closed). The minimum value is one or two orders of magnitude below that which is measured in
conventional wind tunnels; it corresponds to laminar boundary layers on the nozzle walls. The higher
noise levels are caused by radiations from transitional or turbulent wall boundary layers; they are
comparable with the levels reported for other tunnels at the same Mach number. Design and operational
details of new low-disturbance wind tunnels (up to Mach 18) are described in [39].
4.4. further applications of the en method
Transition experiments were carried out in the quiet tunnel: on a flat plate, transition Reynolds numbers as
high as 12 106 were measured. This value is nearly an order of magnitude larger than those obtained in
conventional, noisy facilities [35]. What is interesting to notice is that it corresponds to the theoretical RXT
given by the en method for Me = 3.5 and n = 10 (figure 6).
The flow on supersonic sharp cones constitutes a second case where the free stream Mach number is
constant in the stream wise direction. Measurements performed in the quiet tunnel on a 50 half angle sharp
cone indicated values of RXT close to 7 or 8 106 , two or three times the values obtained in conventional
wind tunnels. The situation is summarized in figure 8 ; it shows flight transition data which were collected
for sharp cones by Beckwith [40]. The transition Reynolds numbers are plotted as a function of the Mach
number. The figure also contains a correlation for wind tunnel transition data, which lies much below the
flight experiments. The range of results obtained in the "quiet tunnel" are reported for comparison.

10

8
64,

RXT

t
10

t.

li"t>t>
t>t>
~t>"t>t>

t.

"tit.

ti

~t.

t.

I At.

%>

"A

t.
<>0

Symbols : free flight conditions


: mean curve for conventional
..nnd tunnels

"quiet tunnel"

t.

10

<>

Me

Figure 8- Transition Reynolds numbers


on cones

0
4
8
12
16
From a theoretical point of view, it can be assumed that the stability equations are the same on cones
and on flat plates, the evolutions of the mean flow properties in both cases being related by Mangler's
transformation. For cones, the predicted transition Reynolds number computed for n = 10 in the quiet
tunnel conditions is 8 106, in very good agreement with experimental data.
The first conclusion is that the en method, with n '" 10, can be applied for predicting transition in
supersonic flow provided the background disturbance level is low enough. Bushnell et al [41] stated that the
applicability of this method is much wider than previously conjectured: n factors of the order of 9 to 11
correlate experiments at low and high speeds and include the effects of Tollmien-Schlichting, Gllrtler and
cross flow instabilities. But what does "low background disturbance level" mean? Is a low turbulence wind
tunnel representative of flight conditions? In this respect, the flight experiments on the AEDC cone
(Fisher and Dougherty, [42]) provided us with some very interesting information. This cone was mounted
on the nose of an F-15 aircraft and flown at Mach numbers from 0.5 to 2 and at altitudes from 1500 m to
15000 m ; the measured transition locations were well correlated with n factors ranging from 9 to 11
(Malik, [43]). However, as it was pointed out by Bushnell et al. [41], in the absence of engine
noise/vibration, the n factors could be of the order of 15, or greater!
The en method was also used by Malik et al [44] for the rather complex reentry-F experiment. The
reentry-F flight vehicle [45] consisted of a 50 semi-vertex cone with an initial nose radius of 2.54 mm.
Stability calculations were performed for an altitude of 30.48 km, where the free-stream Mach number was

Computational Methods in Hypersonic Aerodynamics 245


close to 20. At the measured transition location, the n factor was around 7.5. This somewhat low value
could be due to roughness effects at the junction between the nose and the cone. These computations,
however, extend the e" method into the hypersonic, reacting gas regime.
It must be kept in mind that the en method is an amplification method and not an amplitude method ;
for instance, it cannot take into account the unit Reynolds number effects, which are found to be important
in many experiments. These effects are due in part to a modification of the free stream disturbance level
when the unit Reynolds number is modified and the key problem lies in the understanding of the receptivity
mechanisms, i.e. it is now necessary to establish the link between the forced external disturbances and the
initial amplitude AO of the instability waves. This imposes at first to measure accurately the ambient
disturbances levels (amplitude, spectra, orientations), as well as the mean flow parameters. This latter point
is illustrated in figure 8 : the flight data were obtained for varying conditions of wall temperatures, the
distribution of which is not known in many cases. As the ratio T w rraw strongly affects the stability
properties, there is a large scatter in the data and a quantitative comparison with results deduced from the en
method cannot be made.
5 - SOME PARAMETERS ACTING ON TRANSITION
5.1. Wall temperature effects (flat plate)
The first effect of wall cooling is to modify the evolution of the mean properties of the laminar boundary

rJRx:

layer. For instance, when the wall is cooled, the shape factor H decreases, as well as the ratio R5 1
As
far as the influence of T wrr aw on transition onset is concerned, figure 9 shows experimental results
collected by Potter [46] : the ratio RXT/RxTO is plotted as a function of T wrraw for Mach numbers between
1.4 and 11. RXTO is the transition Reynolds number measured under adiabatic conditions. Although there is
some scatter in the data, it appears that cooling the wall delays transition onset. This effect is rather strong
in the transonic range, but it is greatly reduced when the Mach number increases. Another interesting feature
which can be observed in figure 9 is the appearance of "transition reversals" and "transition re-reversals"
(Me = 3.54 and 8.2). The origin of this behaviour has not been clearly established.
4

_ Cone
--- Flat plate

RXr
RXro

Me = 3.54

8.2

Figure 9- Effect of wall cooling on the transition


Reynolds number

, ' I,\\
/I

"

6.8

~ \ "
6.0,___
4.9
......6.8'::..,..... _
---~ ......

_-

......

......

-""::...':::::::~-~--==-

11.1

0.2

0.4

---==-:--4.7

0.6

....... Tw/Taw
0.8

1.0

The stabili ty of laminar boundary layers on cooled walls was studied by Boehman and Mariscalco [47] in
transonic flow, by Mack [15], Wazzan et al [48], Malik in supersonic flow [49]. Systematic computations
were performed at ONERNCERT by Vignau [50].

246

Computational Methods in Hypersonic Aerodynamics

Ur

t
4

Ur
_ '0

cr;
0

O~5

(b)

2
=3 4

_ 10

(C 3

cr;

0
~e=

C -_-

- - - - -_ _ REi,

o
80000

40000

40~

REi,
800CD

Figure lO-Stability diagrams for Me = 7, '" = 0


Twrr aw = I (a) and 0.3 (b)

Two examples of stability diagrams [50] are given in figure 10 for Me = 7, 'V = 0 (two-dimensional
waves), Twrraw = 1 and 0.3. ar and ai are made dimensionless with the displacement thickness ~1, and
R~1 = Ue~l/Ve. A noticeable feature is that increasing wall cooling tends to separate the second mode
from the first one: for T wrr aw = 0.3, two distinct neutral curves are observed, as it was the case on
adiabatic walls at lower Mach numbers.
Rb1=

Max (-O(i)

30000
26000
17500

0.005

0.05

rnx=150C/
0(1

11)

made dimensionless
with 01

R~1

between

16700 and
33000 )
b)

TwlTaw

Tw/Taw

1~------~O~.5~----~O

Figure II-Maximum value of - <Xi

a., a function of Twrraw

for Me = 7, '" = 0
As T wtraw decreases, the instability of the first mode decreases and vanishes for sufficiently low values
of the cooling mte. As far as the second mode is concerned, figure lla shows the evolution of its
maximum amplification mte as a function of Twtraw for three values of the Reynolds number. It is clear
that the effect of wall cooling is rather small. However, if another reference length is chosen, the
conclusions seem to be different. For instance, the maximum amplification rate which is plotted on
figure lIb was made dimensionless with L = (Vex/Ue)l/2. In this case, it can be said that cooling is
destabilizing. Obviously, when the n factor is computed, the results do not depend on the reference length.

It is interesting to note that the en method (with a fixed value of the n factor) can reproduce the
experimental trends shown on figure 9 : wall cooling has a strong stabilizing effect at low Mach numbers.
This effect is less and less apparent at higher speeds.

--------

Computational Methods in Hypersonic Aerodynamics 247


5.2. Streamwise pressure mdient
The stability res~lts ~el.ated ~ith the infl~e~ce of streamwise pressure gradient at high speed are not
numerous. A major difficulty IS that self-similar boundary layer profiles no longer exist, and it becomes
n~ssary to ~edu~e the basic profiles from non similar boundary layer equations. Zurigat et al (1990, [51]),
for mstance, mvestigated the effects of favourable and adverse pressure gradients for freestream Mach number
distributions of the form Me'" xm. Malik [49] demonstrated that the second mode can be stabilized with
favourable pressure gradients.

a,..

a"

7.

_10 (Xi

=0

0.3

0.9

7.

- 10 (Xi

=0

==--B

la)
I b)

0
0

10000

20000

a
0

10000

20000

Figure 12Stability diagrams for Me = 5.8, 'II = 0


a - Zero pressure gradient
b - Mild positive pressure gradient
c - Strong positive pressure gradient

lei

10000

20000

Vignau [50] analyzed the influence of positive pressure gradients for Me varying linearly with x. The
stability diagrams presented on figure 12 were obtained for the same value of the local freestream Mach
number (Me = 5.8) and three pressure gradients, which are characterized by the value of the incompressible
shape factor Hi : zero pressure gradient (Hi =2.99), mild positive pressure gradient (Hi =3.46) and strong
positive pressure gradient (close to separation, Hi = 4.04). The wall is adiabatic and two-dimensional
waves are considered.
When Hi increases, the first mode is stabilized, but the instability of the second mode becomes more and
more important. A striking feature of the stability diagram associated with the strong pressure gradient is
the appearance of a third unstable mode. Similar evolutions were also computed for Me = 5 and 7 [50].
5.3. Nose bluntness
The effect of nose bluntness on transition has been studied experimentally by many investigators due to its
importance for many hypersonic configurations. Stetson et ai, for instance [52], studied the transition
mechanisms on a cone which could be equipped with interchangeable spherically blunted noses of various
radii. The free stream Mach number was 8 and the half-angle of the cone 7. The main result was that
transition location moved downstream for small values of the nose radius, in agreement with other
experiments (for instance the experiments of Potter and Whitfield [53] on flat plates).

248 Computational Methods in Hypersonic Aerodynamics


The first theoretical attempt at understanding the effect of bluntness was made by Reshotko and Khan
[54] who used the method of multiple scales to analyze the stability of supersonic flows past a blunt flat
plate. More recently, Malik et al [55] performed a linear stability analysis for the experimental conditions
studied by Stetson et al. In order to properly calculate the entropy-layer effects, the basic flow was computed
by solving the Parabolized Navier-Stokes (PNS) equations; the initial conditions for the PNS solution
were provided by solving the full c,Q,...pressible Navier-Stokes equations around the cone nose. Linear
stability analysis of this basic flowvievealM that, with small amount of bluntness, the critical Reynolds
number increases by an order of magnitude compared with the sharp cone value. The computed second mode
frequencies were in reasonable agreement with the experimental results.
By using the en method, it was found that the predicted transition Reynolds number increased due to
small nose bluntness. This increase, however, was much smaller compared to the change in the critical
Reynolds number and depended on the value of the n factor. In fact, the rearward movement of transition can
be partly explained by the strong reduction in local unit Reynolds number associated with the detached
curved shock. The effect of bluntness is to stabilize the initial high frequency disturbances present at small
Reynolds numbers in sharp cone flow.
However, several experiments indicate that the trend is reversed as soon as the nose bluntness exceeds a
certain limit, and the location of transition begins to move forward when bluntness is further increased. For
a large nose radius, Stetson rtrah[52] observed an instability present in the entropy layer; the stability
analysis of Malik et al [55] tOtirlrriled the existence of this low frequency instability between the boundary
layer and the shock, but the computed growth rates were too small to explain the forward movement of
transition, so that the transition reversal for large nose bluntness is not yet fully explained.
5.4. GOrtier instability
For flows developing over a convex surface, centrifugal forces exert a stabilizing effect, in the sense that a
displaced fluid element tends to be restored to its equilibrium position. The magnitude of this effect is
small: in incompressible flow, it was found that, on convex surfaces up to 'Ol/R = 0.0026 (R is the
radius of curvature), the same Tollmien-Schlichting instability occurs as for the flat plate and the transition
Reynolds number remains unchanged.

Figure 13-G6rtler vortices in a flow along a concave wall

On the other hand, the destabilizing effect of centrifugal forces on concave walls leads to the formation
of pairs of counter-rotating vortices, the axes of which are parallel to the principal flow direction
(figure 13). This instability, which was first treated by GOrtier [56], acts in two ways on the boundary layer
development. At first, the GOrtler vortices affect indirectly the transition process by modifying the
development of unstable waves and the breakdown to turbulence occurs earlier than on a flat plate. The
second aspect is essentially important in compressible flows: the vortices induce span wise variations of
the wall heat flux, which result in strong nonhomogeneities of the wall temperature in the laminar regime.
In fact, the appearance of GOrtler instability is not necessarily linked with the presence of a concave
surface; arrays of longitudinal vortices can be observed on flat plates as soon as the model geometry

Computational Methods in Hypersonic Aerodynamics 249


imposes a concave curvature in the streamlines. For example, Ginoux [57] showed evidence of streamwise
vortices at a Mach number of 3, behind a backward facing step. In this case, the boundary layer streamlines,
in separating from the comer and reattaching on the plate, exhibit a strong concave curvature which induces
streamwf~ vortices. Ginoux measured the local heat transfer rates and noted that the effect of these vortices
was to poduce locally very large peaks in the heat rate, much larger than the usually measured turbulent
values inlmediately after transition.
Stream wise vortices were also observed by many authors behind a wire roughness or downstream the
reattachment line in a compression comer (Delery-Coet, [58]). In all cases, the vortices are responsible for
large spanwise variations of the wall heat flux.
On the theoretical side, the problem of GOrtler vortices can be treated by using a linearized stability
theory. Several papers were devoted to stability analyses in incompressible flow (see [59], for example), but
calculations including compressibility effects are not numerous. EI Hady and Verma [60] investigated the
growth of stream wise vortices in two-dimensional boundary layers along curved surfaces over a range of
Mach numbers from 0 to 5; similar computations were performed at ONERA/CERT by lallade [61].
These calculations use a curvilinear system of coordinates representing streamlines and potential lines of the
inviscid flow. It can be demonstrated that the mean flow is described at the first order by the conventional
compressible boundary layer equations. A steady three-dimensional small disturbance is then superposed on
this two-dimensional basic flow ; the disturbances under consideration are assumed in the form :
r' = r (y) cos(az) exp(crx), with r' = u', v', p' or T'
w'

=w

(12)
(y) sin(az) exp(crx)

where a is the wavenumber in the spanwise direction and cr the spatial growth rate in the x-direction along
the curved wall. a and cr are usually made dimensionless with the characteristic length L = (vooX/uoo)1/2.
As for the Tollmien-Schlichting waves problem, relations (12) are substituted into the linearized NavierStokes equations. Keeping the leading term leads to an eighth order system of homogeneous, linear,
ordinary differential equations with homogeneous boundary conditions. This forms an eigenvalue problem
for the three real parameters a, cr and G. G is the Gortler number, defined as :
G = uooL
Voo

(1.)
1/2
R

(13)

where R is the radius of curvature of the wall.

100

100

0=

0= 10
5

20

10

10
0.1

(al
0.1
0.01

0,1

(b 1

-Q
01 0.01
Figure 14-Stability diagrams

-Q
0.1

250

Computational Methods in Hypersonic Aerodynamics

Two stability diagrams are presented in figure 14 for an adiabatic wall and two Mach numbers (Moo = 0
and 3). Curves of constant amplification rate are plotted in the G()rtier number-wavenumber plane. The
shape of these curves does not change very much when the Mach number increases; as for the TollmienSchlichting waves, there is a critical G()rtler number, G cr , below which the flow is stable for any
disturbance wavenumber. It can be observed that compressibility has a stabilizing effect: as Moo increases,
the maximum growth rate at a given G()rtler number decreases; however, the influence of compressibility
on reducing cr becomes very small at high values of G. Computations by EI Hady and Verma [62] showed
also that G()rtier instability is more difficult to influence and control by suction or cooling than TollmienSchlichting instability. Malik [23] demonstrated that real gas has very little effect on the G()rtier vortices.
It is then possible to calculate the total amplification rate InNAo by integrating the local growth rate in
the stream wise direction. For the Tollmien-Schlichting waves, this integration is carried out for each
frequency; for the G()rtler vortices, it is performed for each spanwise wavelength. This was done by EI
Hady and Verma, who concluded that compressibility reduces the maximum amplitude ratio by 20 % as
Mach number increases from 0 to 5. Beckwith et al [63] used the en method in order to predict transition
location along the nozzle walls of the "quiet tunnel" ; they found that Tollmien-Schlichting instability is
weak, and that transition results in the growth of G()rtier vortices along the concave walls. Nevertheless,
calibration of the en method for the G()rtier instability remains an open question.

Quite recently, a new approach has been developed to study the linear stability problems for TS waves
(Herbert [64]) as well as for G()rtier vortices (Spall and Malik [65]). In this approach, it is now assumed
that the amplitude functions of the disturbances (which depend on y only in the framework of the classical
linear theories) depend also on the streamwise distance x. For the G()rtier vortices, the term exp(crx)
disappears, and the disturbances are written as :
r' = r (x, y) cos(az) with r' = u', v', p', or T'
(14)
w' =w (x, y) sin(az)
Substituting the previous expression into the stability equations leads to a system of partial differential
equations of parabolic type which can be solved using a marching procedure in x (pSE : Parabolic Stability
Equations). This approach makes it possible to take into account the non parallel effects.
Spall and Malik [65] numerically studied the development of G()rtler vortices in supersonic and
hypersonic boundary layers by solving the parabolic partial differential equations. They determined the effect
of various initial conditions on G()rtler instability and computed growth rates which were somewhat
different from those given by a normal mode solution, although the trends were identical.
5.5. Bypass by large roughness elements
It has been shown in the previous paragraphs that the linear stability theory, associated with the en method,
can describe some of the essential features of the transition process when "natural" instability waves are
present However, these waves no longer appear if the amplitude of the forced disturbances is too large, so
that the problem becomes fully non linear: it is a "bypass" (Morkovin [66]), in the sense that the linear
mechanisms are completely ignored. As there is no general bypass theory, it is necessary to develop
different criteria for each type of bypass (the word criterion must be interpreted as an empirical correlation
between boundary layer and flow parameters at transition onset).
Boundary layer tripping by large roughness elements constitutes a typical example of bypass. When
isolated tripping devices (spheres, small cylinders normal to the wall, ... ) are introduced into a nominally
two-dimensional boundary layer developing on a flat plate, wall visualisations show the existence of
horseshoe vortices wrapped around the roughness element, see figure 15. At a distance L from the tripping
device, the "legs" of these vortices break down and create a turbulent wedge. When the height k of the
protuberance increases, L decreases up to a minimum value which is reached for an "effective" roughness
size, kerr. In other words, L remains constant for k > kerr.

-----------

Computational Methods in Hypersonic Aerodynamics

vortices

251

"tur bulent"
wedge

t
_x
Figure IS-Example of wall visualization using thermosensitive paint

From a practical point of view, the value of kerr is of great importance. Van Driest and Blumer ([67] to
[69]) performed a series of experiments in order to deduce empirical correlations between Rkeff, RXk and
the flow parameters; Rkeff and RXk are the Reynolds numbers formed with the effective roughness height
and the roughness location, respectively. The measurements were made on cones and correlated by the
following relationship:
Rkeff = 33.4

[1 + Y

M~ - 0.81 r-awT~ Tw)] Rx~4

(15)

Van Driest and Blumer assumed that this expression was also applicable to flat plate flows by using
Mangler's transformation, which simply consists in replacing the coefficient 33.4 by 33.4 (3)1/4 = 44.
Vignau [50] demonstrated that the modified correlation largely underestimates the effective roughness height
for flat plate flows, because Mangler's transformation is valid for mean flow properties, but it cannot be
used for stability and transition problems.

6 - TRANSITION REGION (TWO-DIMENSIONAL FLAT PLATE FLOW)


6.1. Definition of the transition region
Transition starts when the first turbulent structures (spots) appear in the laminar boundary layer. In natural
conditions, the spots originate in a more or less random fashion. Once created, they are swept along with
the mean flow, growing laterally and axially and finally covering the entire surface. The transition region is
defined as the region where the spots grow, overlap and form a turbulent boundary layer. When a hot wire is
placed in the boundary layer (or when a film gage is mounted flush with the model surface), the fluctuations
which are recorded in the transition region show the successive appearance of turbulent spots and of laminar
regions; it is the intermittency phenomenon. The intermittency factor y represents the fraction of the
total time that the flow is turbulent. Experimentally, it is not always easy to define the beginning (y = 0)
and the end (y = 1) of transition.
The beginning of transition is often taken at the point of initial measurable deviation of a characteristic
parameter from its laminar evolution. This can be, for instance, the beginning of a faster growth of the
boundary layer thickness, the point of minimum surface temperature (or surface Pitot pressure or wall heat
flux), the location where the increase in the rms voltage from a hot wire becomes steeper ... As it has been
pointed out by Owen and Horstman [70] most of the transition data reported for high speed flows are not
based on direct observations of turbulent spots, but rather on the evolution of some macroscopic parameter
(skin friction, heat transfer, ...) ; their departure from laminar values can be detected only when the
intermittency is appreciably greater than zero.

252 Computational Methods in Hypersonic Aerodynamics


In the transition region, it is well known that quantities such as nns voltage, skin friction, wall heat
flux, ... reach a maximum at nearly the same location [71] and then decrease more or less slowly. The peak
value of these quantities is often used to define the end of transition, because its position is easy to measure
accurately. However, these points are located upstream of the end of transition. For instance, it is
established (Owen, [71]) "that the peak nns signal coincides with the point where the turbulent burst
frequency is maximum" and not with the point where the boundary layer is fully turbulent.
These observations can explain to a great extent the large scatter which is observed in transition data.
Inconsistent choices of criteria for the beginning and the end of transition make it difficult to compare
experimental results obtained through different techniques. However, in spite of these problems, some
general trends have been put forward, as it will be shown below.
6.2. Extent of the transition re!1;ion
Let us assume now that the locations of transition onset, xT, and of transition end, xE, are measured in a
consistent way. ~x = xE - xT represents the transition extent and the Reynolds numbers RXT, RXE, R~x
are based on xT, xE and ~x, respectively. The evolution of R~x as a function of RXE is given in figure 16
for free stream Mach numbers between 0 and 8 ; these data were obtained or collected by Potter and
Whitfield [53], who defined XT and XE by examining the boundary layer growth from schlieren
photographs. At a given value of RXE, it is clear that a significant increase in R~x is associated with
increasing Mach number. Chen and Thyson [72] suggested the following relationship:

R~x

= (60 + 4.86

M~92) RX~

(16)

Me =

8
Figure 16-Extent of the transition region
which reflects the experimental trend. Parameters such as wall temperature, unit Reynolds number, ... , are
not taken into account in these correlations. It was noted by Morkovin [73] that "the lateral or transverse
growth of a turbulent spot decreases from about 11 0 semi-angle at low speeds to about half the angle at
hypersonic speeds". The same observation can be made for the spreading of a turbulent wedge behind an
isolated roughness element. Both phenomena involve the same physical mechanism which is called
"transverse contamination".
6.3. Transition region modellin!1;
From a practical point of view, the modelling of the transition region becomes a more and more important
problem when the Mach number increases; this is due to the fact that, for supersonic and hypersonic
flows, the evolution from the laminar to the turbulent state occurs along a stream wise distance which can
be much more important than the laminar region extent which precedes it.

Computational Methods in Hypersonic Aerodynamics


.

253

A prac~ical calculation method was developed at ONERA/CERT for the two-dimensional,

Incompressl~le flows [74] and then extended to high speed conditions: it is assumed that the turbulent

shear stress IS expressed by :

- p u'v' = E Ilt au/ay

(17)

Ilt is an eddy viscosity coefficient, which is computed by using a classical turbulence model. At first
sight, the ~oefficient E represents the intermittency factor y, in the sense that it increases from 0 in laminar
flow to 1 In turbulent flow.

1.5 r---,.-=-------~

1.01--/--------;:::::".---1

Experiments: Juillen
Fla t

Figure I7-"Intermittency" function

pia te

relations

Calculations

OL-----~----~

____

0.2

-X(m)

- L_ _ _ _~_ _ _ _ _ __

0.4

Figure I8-Comparison with experimental results


obtained by Juillen
From experimental data at low s.peeds, it was found that the momentum thickness 9E at the completion
of transition was about twice the momentum thickness 9T at the transition onset. This led to the idea that E
could be represented as a function of 9/9T. This function was determined to fit available experimental
results in zero and positive pressure gradients. It is shown in figure 17 ; one of the problems was to model
the overshoot in the skin friction coefficient, which exists in the middle of the transition region due to the
intermittency phenomenon. This was done by imposing an overshoot to the function E, which does not
represent the physical intermittency factor, but rather an empirical weighting coefficient for the Reynolds
shear stress.
This method was extended to supersonic and hypersonic flows by taking into account the lengthening of
the transition region at high speeds. For this, the analytical expression of the function E remains the same,
the parameter 9/9T - 1 being simply replaced by (9/9T -1)/(1 + 0.02 M;).
A first example of application is given in figure 18. The calculated wall heat flux is compared with
measurements performed by Juillen [36] on a flat plate for Me = 7. The location of transition onset is
imposed (xT = 0.1 m). Laminar and turbulent curves deduced from analytical flat plate relations are also
shown. The "intermittency" method gives good results; in particular, the maximum of the heat flux is
well predicted. This maximum represents an overshoot above the fully turbulent value, as it can be seen by
comparison with the turbulent curve; it has been noticed previously that this point is associated with the
maximum spots frequency and not with the completion of transition. In fact, the fully turbulent properties
are only achieved towards the end of the plate.

254

Computational Methods in Hypersonic Aerodynamics


5

10- Re/cm
o
~

} Experiments

Coleman, Elfstrom and Stollery

Calcula tions

(w/(m~~)

5.5

2.76
1.46

____________________________- .
Ti " 1070 K
Tw = 295 K

8
6

4
2
O~------L-------L-------~----~

~x(m)

0.6

0.4

0.2

Figure 19-Comparison with experimental result" obtained by Coleman et at.


Figure 19 shows another example of comparison between experiments and calculations. The
measurements were carried out by Coleman et al.[75] at a free stream Mach number equal to 9 and for three
values of the unit Reynolds number. As for the previous case, the position of transition onset is given in
the computations. The increase in the wall heat flux and the location of its maximum are well predicted, but
large discrepancies are observed further downstream. The calculations indicate that the transition region
extends up to x'" 0.6 m.
/1/1.

- GJ Pt:O 92 compu tations


-- lZl Pt" 0.89
GJ

0.90
0.85

Experiments IChen et at.)

,)"""m",'

\'I/'!.":j!..j~~!..!.'..!!..'.!.'!.!
X(m)

0,80
0,15

0,20

0.25

0,15

0,20

0,25

X(m)

Figure 20-Transition in the "quiet" tunnel (Chen et al)


A last comparison is given in figure 20, which is relative to the experiments performed in the quiet
tunnel on a 50 half angle cone (Chen et ai, [35]). The wall is adiabatic, the free stream Mach number is 3.5,
the unit Reynolds number and the transition Reynolds number are 49,2 106 m- l and 8.46106 m- l ,

Computational Methods in Hypersonic Aerodynamics 255


respectively. In the computations, the transition location is imposed, and it is assumed that the turbulent
Prandtl number Pt is constant. Two values of Pt were used: 0.89 and 0.92. Figure 20 presents the
theoretical evolution of the skin friction coefficient, as well as a comparison between measured and
calculated distributions of the recovery factor r = (Taw - T e)/(Tie - Td. It appears that the skin friction
and the extent of the transition region do not depend on the value of Pt ; but a better agreement with the
experimental evolution of r is achieved with Pt = 0.92. If ~x is defined as the distance between the
minimum and the maximum values of Cf, we observe that the ratio ~x/xT = RAx/RxT is close to 0.3,
whereas, in the previous two examples, this ratio was about three times larger. This is the result of two
combined effects: at first, the Mach number is smaller than in the previous cases, and it is established that
increasing Me increases the ratio R~x/RxT, see for instance relation (16). A second parameter is the
transition Reynolds number: it is very large in the quiet tunnel experiments, and relation (16) indicates
that R~x/RxT varies as Rxi/3. As the experiments reported in figures 18 and 19 were performed in noisy
wind tunnels. the corresponding values of RXT are rather small and the question arises of the extent of the
transition region at hypersonic speeds in a low disturbance environment, i.e. at very large values of RXT
like those predicted by the en method with n "" 10 : if the trends given by (16) are correct, the problem of
transition region modelling could become less crucial.
Malik [44] reported other applications of the previously discussed intermittency model. He stated that
"the error in the extent of the transition region was of the same order as the error in the prediction of the
transition onset location by the en method".
7 - THREE-DIMENSIONAL BASIC FLOWS
7 .1. Cross flow instability
When a boundary layer develops on a three-dimensional geometry, such as a swept wing, a swept cylinder
or a body at incidence, the mean velocity profile becomes twisted. It is usually decomposed into a
streamwise profile u (in the direction of the external streamline) and a crossflow profile w (in the direction
normal to this streamline), as iIIustrated in figure 21.
y

wall
streamline

Figure 2 I-Stream wise and crosstlow mean velocity profiles


In incompressible flow, the mechanisms of three-dimensional transition are relatively well understood,
see review papers by Poll [76], Amal [77], Saric and Reed [78]. As a first approximation, it can be assumed
that transition is induced either by streamwise instability or by crossflow instability. The streamwise mean
velocity profiles look like classical two-dimensional profiles; they are essentially unstable in positive
pressure gradients, where they induce transition through the action of an inflectional instability. On the
other side, an inflection point is always present in the crossflow mean velocity profiles. As these profiles
develop rapidly in regions of strong negative pressure gradients, transitions of the crossflow type are
expected to occur in accelerated flows, for instance in the vicinity of the leading edge of a swept wing. In
addition, a linear stability analysis shows that crossflow instability can amplify zero frequency

256

Computational Methods in Hypersonic Aerodynamics

disturbances; this leads to the formation of stationary, corotating vortices aligned in the local streamwise
direction. In the experiments, crossflow vortices are observed as regularly spaced streaks.
Experimental and numerical results on crossflow instability are not numerous for supersonic and
hypersonic flows. Balakumar and Reed [79] performed stability computations for a supersonic boundary
layer developing on a rotating cone at zero angle of attack. Transition on a cone at angle of attack was
investigated by Marcillat at Moo = 5 [80] and by Stetson et aI at Moo = 8 [81] ; Creel et aI [82] reported
measurements on circular cylinders at sweep angles of 45 and 60 ; the data were obtained in the Mach 3.5
"quiet tunnel" at NASA LANGLEY. In the following paragraph, the problem of crossflow instability is
illustrated by experiments performed in the supersonic R3Ch wind tunnel of the ONERA CHALAIS
MEUOON Center [83].
7.2. Example of theoretical and experimental results
The R3Ch wind tunnel is a free jet, blowdown wind tunnel with a circular nozzle, the exit diameter of
which is 34 cm. The freestream Mach number Moo is 10, for a stagnation temperature and a stagnation
pressure equal to 1 100 K and 120 bars, respectively (unit Reynolds number close to 10 106 m l ).
Static pressure
+ Thermocouples

Z(m)

Figure 22-Experimental setup

Figure 23-Thermosensitive paint visualization

The model is a swept circular cylinder (diameter D = 6 cm) equipped with an hemispherical nose, see
figure 22. In the present experiments, the sweep angle cp is equal to 40. X and Z represent the directions
normal and parallel to the attachment line (and to the cylinder axis), respectively. Let us recall that the
attachment line is the line where the static pressure is maximum on an infinite swept body of constant
chord. As X is measured along the surface of the model, the azimuthal angle 9 is equal to 2X/D. The
attachment line corresponds to X =0, i.e. 9 =0. In fact two models were used: one is made of silastene,
in order to perform wall visualizations by using a thermosensitive paint; the second one is made of steel
and it is equipped with a row of thermocouples and a row of static pressure taps distributed in the span wise
direction Z, see figure 22.
Figure 23 shows a wall visualization obtained with the first model. For 9 = 40, where the flow is
strongly accelerated, one can observe streaks which are practically aligned with the external streamline.
These streaks constitute the signature of the cross flow instability, which amplifies zero frequency,
stationary waves. The linear stability theory was used in order to see if it was possible to predict the
wavelength of the streaks. The laminar boundary layer was first computed from the measured pressure
distribution, which closely agreed with the newtonian law. The mean velocity and mean temperature
profiles obtained at 9 = 40 were then introduced as basic profiles in the stability code. It was assumed that
the disturbances were expressed by :
r' = r(y) exp(- aiX) exp [i(arX + PrZ)]

(18)

Computational Methods in Hypersonic Aerodynamics 257


i.e. there is no amplification in the span wise direction.
0,5

A(mm)::2n/(a~ +f3~)1!2
2

Ej
Theory

ljJ(O)
~--~~~~~~~~

500
o~~

...... Tw{K)
______
~______~

300

400

500

Figure 25-Wavelength of the streaks


Figure 24-Theoretical amplification rates
Figure 24 shows the evolution of the spatial amplification rate (Xi as a function of the 'I' angle (angle
between the external streamline and the wavenumber vector), for three values of the wall temperature:
TW = 290 K (temperature at the beginning of the run), 350 K and 500 K. The latter is approximately the
wall temperature of the model at the time where the photograph of figure 23 was taken. It is clear that
increasing TW has a destabilizing effect. For Tw = 500 K, one can observe a narrow range of unstable
directions, between 84.8 and 85.4. This confirms that the cross flow instability amplifies the stationary
disturbances as soon as the wall temperature becomes high enough. Figure 25 presents the variation of the
wavelength of the less stable (Tw < 350 K) or of the most unstable (Tw ~ 350 K) stationary
disturbances, as a function of the wall temperature. For TW '" 500 K, the computations are in good
agreement with the measurements. A similar agreement was reported in [84] for experiments performed in
the "quiet tunnel" at NASA Langley (Moo = 3.5).
The n factor was also computed for several zero and non zero unstable frequencies between 9 = 0 and
9 = 80. Its maximum value is lower than 2. Experimentally, no "natural" transition was detected on the
cylinder. In fact, the validity of the en method is not yet established for three-dimensional, supersonic
flows, even if it seems to work at lower speeds for predicting crossflow-induced transitions.
7.3. Leading edge contamination
When a swept body is in contact with a solid wall (fuselage, wind tunnel wall ...), it has been observed that
the large turbulent structures coming from the wall at which the model is fixed may develop along the
attachment line: it is the so-called leading edge contamination. A similar phenomenon is likely to occur
when large isolated roughness elements are placed on the attachment line.
In incompressible flow, the leading edge contamination criterion, first proposed by Pfenninger [85] is
based on the value of a Reynolds number R defined as :
R

=WeTlIv

(19)

258

Computational Methods in Hypersonic Aerodynamics

1\

1\

with 11 = (v/klfl.) and k=

[dU]
~ X::::O

U e and We are the freestream velocity components in the X and Z directions. For R > 245, leading
edge contamination occurs: the turbulent structures generated by a source of gross disturbances become
self-sustaining and develop in the spanwise direction. They are damped and disappear for R < 245. Poll
[86] extended this criterion to compressible flows by introducing a transfonned Reynolds number R

which has the same definition as R , except that the kinematic viscosity v is replaced by v* which is
computed at a reference temperature T*.

O.

O~.

D=4cm
D=6cm

400

Tripping
00 No tripping
e Not clear
(intermittent?)

..

200

POLL

0 0 0

Figure 26-Wall heat flux distribution

Figure 27-Leading edge contamination criterion

Leading edge contamination experiments were carried out by placing roughness elements on the
attachment line of swept cylinders in the R3Ch wind tunnel (paragraph 7.2.). As an example of result,
figure 26 shows a three-dimensional picture of the distribution of the wall heat flux coefficient h for a
cylinder of diameter D = 6 cm and an angle of sweep cp = 40 ; this corresponds to R * = 326. The
spanwise position is given by the thennocouple number, see figure 22. The roughness element is a small
cylinder nonnal to the wall; its height k is equal to its diameter; in the present case, k = 0.6 mm, about
two times the boundary layer thickness. An important increase in h is observed around the tripping device,
which constitutes the apex of a turbulent wedge developing with a spreading half angle close to 15. At the
last thennocouple position, the boundary layer is fully turbulent.
Similar experiments were carried out for several values of D, cp and k. The results are summarized on
-

1\

figure 27 in the ( R *, k/11 *) plane, with 11 * = (v* / k) 1/2. For large values of

k/T1 *, it appears that the

tripping becomes effective as soon as R * exceeds a critical value close to 250, in agreement with Poll's
criterion (see also Da Costa and Alziary [87]).
Additional measurements were perfonned by placing the roughness element at non zero values of Elk: (9k
denotes the azimuthal angle of the device location). It was found that the minimum roughness height

Computational Methods in Hypersonic Aerodynamics 259


which is necessary to trigger transition increases rapidly with 9I<: [83]. as conjectured by Morrisette [88] and
Poll [86].
8 - CONCLUSIONS
It is clear that, even in the "simple" case of two-dimensional mean flows, the problems associated with
boundary layer transition at high speeds are numerous and that many of them are far from being solved.

On the theoretical point of view, the linear stability theory constitutes a very efficient tool to understand
the fundamental mechanisms leading to transition and sophisticated experimental studies gradually confmn
these theoretical elements. The linear theory can also explain, at least qualitatively, the influence of more or
less controlled parameters (bluntness, wall cooling for instance), but the key problem lies in the
understanding of the receptivity mechanisms and the exact relation between instability and transition is not
very well known. Despite these facts, the extension of the en method in supersonic and hypersonic flow
could provide a reliable prediction of transition; for a low disturbance environment, the n factor has
roughly the same value as in incompressible flow.
As far as the transition region is concerned, we believe that simple methods, such as the intermittency
method, are able to give right predictions in very different situations. The inaccuracies of these techniques
are certainly small as compared with large errors which can arise in the prediction of the transition onset.
As examples of bypass-induced transitions, the effects of isolated, three-dimensional roughness elements
have been studied. For two-dimensional mean flows, the first effect of these tripping devices is to generate
turbulent wedges, the apexes of which move upstream when the roughness size increases. As this movement
cannot be described by the linear stability theory, purely empirical criteria need to be developed. The second
effect is the appearance of horseshoe vortices around the roughness element; these vortices are responsible
for very large values of the wall heat flux, as it is also observed for Gl>rtler vortices as well as for crossflow
vortices; in fact, the problem of minimizing the heat transfer rates in hypersonic boundary layers requires
not only to delay transition, but also to avoid stream wise vortices in the laminar boundary layer. For threedimensional mean flows, it has been demonstrated that Poll's criterion is valid at high speeds, but that its
validity is restricted to the attachment line, which is the location where the boundary layer is the most
sensitive to roughness elements. At the present time unfortunately, there is no general criterion which could
be able to predict boundary layer tripping in a broad range of practical conditions.
REFERENCES

1.

Morkovin, M.V. Critical evaluation of transition from laminar to turbulent shear layers with
emphasis on hypersonically travelling bodies, Report AFFDL-TR-68-149, Wright-Patterson Air Force
Base, Ohio, 1968

2.

Morkovin, M.V. Transition at hypersonic speeds, ICASE NASA Contractor Report 178 315,1987

3.

Reshotko, E. Boundary layer stability and transition, Annual Review in Fluid Mechanics, Vol. 8,
1976

4.

Mack, L.M. Boundary layer linear stability theory, AGARD Report nO 709,1984

5.

Pate, S.R. Effects of wind tunnel disturbances on boundary layer transition with emphasis on radiated
noise: a review, AIAA Paper 80-0431, Colorado Springs, 1980

6.

Amal, D. Laminar-turbulent transition problems in supersonic and hypersonic flows, AGARD Report
nO 761, 1988

7.

Stetson, K.F. Hypersonic boundary layer transition, 3rd Joint Europe/US Short Course in
Hypersonics, Aachen, 1990

260 Computational Methods in Hypersonic Aerodynamics


8.

Schubauer, G.B. and Skramstad, H.K. Laminar boundary layer oscillations and transition on a flat
plate, Report 909 NACA, 1948

9.

Schlichting, H. Boundary layer theory, 6th ed. McGraw Hill, New York, 1968

10. Lees, L. and Lin, C.C. Investigation of the stability of the laminar boundary layer in a compressible
fluid, NACA TN W 1115, 1946
11. Lees, L. The stability of the laminar boundary layer in a compressible fluid, NACA TN W 876, 1947
12. Dunn, D.W. and Lin, C.C. On the stability of the laminar boundary layer in a compressible fluid, J.
aero. Sci., Vol. 22, pp. 455-477, 1955

13. Lees, L. and Reshotko, E. Stability of the compressible laminar boundary layer, J.F.M., Vol. 12, Part
4,pp.555-590, 1962
14. Mack, L.M. Stability of the compressible laminar boundary layer according to a direct numerical
solution, AGARDograph 97, Part I, pp. 329-362, 1965
15. Mack, L.M. Boundary layer stability theory (2 volumes), Jet Propulsion Laboratory, California Inst.
of Techn., Pasadena, California, November 1969
16. Mack, L.M. Linear stability and the problem of supersonic boundary layer transition, AIAA J., Vol.
13, N 3, pp. 278-289, 1975
17. Mack, L.M. Transition prediction and linear stability theory, AGARD Conf. Proc. nO 224, Paris,
1977
18. Mack, L.M. Compressible boundary layer stability calculations for sweptback wings with suction,
AIAA J., Vol. 20, pp. 363-369, 1981
19. Mack, L.M. Remarks on disputed numerical results in compressible boundary layer stability theory,
Phys. Fluids 27(2), February 1984
20. Gaster, M. A note on the relation between temporally increasing and spatially increasing disturbances
in hydrodynamic stability, J.F.M., Vol. 14, pp. 222-224, 1962
21. Arnal, D. Stabilite et transition des couches limites laminaires bidimensionnelles en ecoulement
compressible, sur paroi athermane, La Recherche Aerospatiale nO 1988-4, 1988
22. Malik, M.R. Transition in hypersonic boundary layers, Numerical aJld Physical Aspects of
Aerodynamic Flows IV, Springer Verlag, T. Cebeci ed., 1990
23. Malik, M.R. Stability theory for chemically reacting flows, IUTAM Symp. "Laminar-Turbulent
Transition", Springer Verlag, D. Amal and R. Michel ed., 1990
24. Gasperas, G. Stability of the laminar boundary layer for an imperfect gas, IUTAM Symp. "LaminarTurbulent Transition", Springer Verlag, D. Amal and R. Michel ed., 1990
25. Stuckert, G.K. and Reed, H.L. Stability of hypersonic, chemically reacting viscous flows, AIAA
Paper 90-1529,1990
26. Goldstein, M.E. The evolution of TS waves near a leading edge, J.F.M. Vol. 127, pp. 59-81, 1983
27. Kerschen, E.J. Boundary layer receptivity, AIAA Paper 89-1109,1989

Computational Methods in Hypersonic Aerodynamics 261


28. Choudhari, M. and Strett, C.L. A finite Reynolds number approach for receptivity due to rapid
variations in surface boundary conditions, 1st ASME/ISME Fluids Eng. Conf., Portland, D.C. Reda,
H.L. Reed, R. Kobayashi ed., 1991
29. Erlebacher, G. and Hussaini, M.Y. Second mode interactions in supersonic boundary layers, IUTAM
Symp. "Laminar-Turbulent Transition", Springer Verlag, D. Amal and R. Michel ed., 1990

30. Normand, X. Transition la turbulence dans les ecoulements cisailles compressibles Iibres ou
parietaux, Grenoble University, Thesis. 1990
31. Thumm, A, Wolz, W. and Fasel, H. Numerical simulation of spatially growing three-dimensional
disturbance waves in compressible boundary layers, IUTAM Symp. "Laminar-Turbulent Transition",
Springer Verlag, D. Amal and R. Michel ed., 1990
32. Masad, I.A. and Nayfeh, A.H. Subharmonic instability of compressible boundary layers, Phys. Fluids
A, Vol. 2, nO 8, pp. 1380-1392, 1990
33. Smith, A.M.O. and Gamberoni, N. Transition, pressure gradient and stability theory, Douglas Aircraft
Co. Rept ES 26388, EI Segundo, California, 1956
34. Van Ingen, J.L. A suggested semi-empirical method for the calculation of the boundary layer transition
region, Univ. of Techn., Dept. of Aero. Eng., Rept. UTH-74 Delft, 1956
35. Chen, F. J. and Malik, M. R. Comparison of boundary layer transition on a cone and flat plate at
Mach 3.5, AIAA Paper 88-0411,1988
36. Juillen, J.C. Determination experimentale de la region de transition sur une plaque plane
et 7, ONERA Technical Report W 10/2334 AN, 1969

a M = 5,6

37. Harvey, W.D. Influence of free stream disturbances on boundary layer transition, NASA Technical
Memorandum 78635,1978
38. Beckwith, I. E., Creel Jr, T. R., Chen, F. J. and KendaIl 1. M. Freestream noise and transition
measurements on a cone in a Mach 3.5 pilot low-disturbance tunnel, NASA TP 2180,1983
39. Beckwith, I.E., Chen, F., Wilkinson, S., Malik, M. and Tuttle, D. Design and operational features of
low-disturbance wind tunnels at NASA Langley for Mach numbers from 3.5 to 18, AIAA Paper 901391, 1990
40. Beckwith, I.E. Development of a high Reynolds number quiet tunnel for transition research, AIAA
Journal, Vol. 13, W 3,1975
41. BushneIl, D.M., Malik, M.R. and Harvey W.D. Transition prediction in external flows via linear
stability theory, IUTAM Symp. Transsonicum III, G&tingen, Germany, May 24-27,1988
42. Fisher, D.F. and Dougherty, N.S. In-flight transition measurements on a 10 cone at Mach numbers
from 0.5 to 2, NASA TP 1971, 1982
43. Malik, M.R. Instability and transition in supersonic boundary layers in Laminar-turbulent boundary
layers (E.M. Uram and H.E. Weber Ed.) - Proc. of Energy Resources Technology Conference, New
Orleans, Louisiana, 1984
44. Malik, M.R. Boundary layer transition in hypersonic flows, AIAA Paper 90-5232,1990
45. Johnson, C.B., Stainback, P.C., Wiker, K.C. and Bony, L.R. Boundary layer edge conditions and
transition Reynolds number data for a flight test at Mach 20 (Reentry F), NASA TM X-2584, 1972

262 Computational Methods in Hypersonic Aerodynamics


46. Potter, J.L. Review of the influence of cooled walls on boundary layer transition, AlAA Journal, Vol.
18, n 8, 1980
47. Boehman, L.I. and Mariscalco, M.G. The stability of highly cooled compressible laminar boundary
layer, AFFDL-TR-76-148, 1976
48. Wazzan, A.R. and Taghavi, H. The effect of heat transfer on three-dimensional spatial stability and
transition on flat plate boundary layer at Mach 3, Int. J. Heat Mass Transfer, Vol. 25, nO 9, pp. 13211331,1982
49. Malik, M.R. Prediction and control of transition in hypersonic boundary layers, AlAA Paper nO 871414, June 1987
50. Vignau, F. Etude theorique et experimentale de la transition en ecoulement bidimensionnel
compressible, Thesis, ENSAE Toulouse, 1989
51. Zurigat, Y.H., Nayfeh, A.H. and Masad J .A. Effect of pressure gradient on the stability of
compressible boundary layers, AlAA Paper 90-1451,1990
52. Stetson, K.F., Thompson, E.R., Donaldson, J.C. and Siler, L.G. Laminar boundary layer stability
experiments on a cone at Mach 8 - Part 2: blunt cone, AIAA Paper n 84-0006, 1984
53. Potter, J.L. and Whitfield, J.D. Effects of slight nose bluntness and roughness on boundary layer
transition in supersonic flows, J.F.M., Vol. 12, Part 4, pp. 501-535, 1962
54. Reshotko, E. and Khan, M.M.S. Stability of the laminar boundary layer on a blunted plate in
supersonic flow, 1st IUTAM Symp. "Laminar-Turbulent Transition", Springer Verlag, 1979
55. Malik, M. R., Spall, R. E. and Chang, C. L. Effect of nose bluntness on boundary layer stability and
transition, AlAA Paper 90-0112, 1990
56. GOrtler, H. On the three-dimensional instability of laminar boundary layers on concave walls, NACA
T.M. 1375, 1954
57. Ginoux, JJ. Streamwise vortices in laminar flow. Recent Developments in Boundary Layer Research,
AGARDograph 97, Part I, pp. 395-422, 1965
58. Delery, J. and Coet, M.C. Shock-shock and shock wave-boundary layer interactions in hypersonic
flows, AGARD VKI Special Course on Aerothermodynamics of Hypersonic Vehicles, 1988
59. Floryan, J.M. and Saric, W.S. Stability of GOrtler vortices in boundary layers, AIAA Journal, Vol.
20, N 3, 1982
60. EI Hady, N.M. and Verma, A.K. Growth of GOrtler vortices in compressible boundary layers along
curved surfaces, J. of Eng. and App. Sci., Vol. 2, pp. 213-238,1983
61. Jallade, S. Etude theorique et numerique de l'instabilite de GOrtler, Thesis, Institut National
Polytechnique de Toulouse, 1990
62. EI Hady, N.M. and Verma, A.K. ,Instability of compressible boundary layers along curved walls with
suction or cooling, AIAA Journal, Vol. 22, N 2,1984
63

Beckwith, J.E., Malik, M.R., Chen, FJ. and Bushnell, D.M. Effects of nozzle design parameters on
the extent of quiet test flow at Mach 3.5, IUTAM Symp. on Laminar-Turbulent Transition,
Novosibirsk, Springer Verlag, Ed. V.V. Kozlov, 1984

Computational Methods in Hypersonic Aerodynamics 263


64. Herbert, T. Boundary layer transition - analysis and prediction revisited, AIAA Paper 91-0737, 1991
65. Spall, R.E. and Malik, M.R. G6rt1er vortices in supersonic and hypersonic boundary layers, Phys.
Fluids Al(ll), pp. 1822-1835, 1989
66. Morkovin, M.V. Bypass transition to turbulence and research desiderata, Symp. "Transition in
Turbines", Cleveland, Ohio, 1984
67. Van Driest, E. R. and Blumer, C. B. Boundary layer transition at supersonic speeds - Threedimensional roughness effects (spheres), J. of the Aer. Sci., Vol. 29, W 8,1962
68. Van Driest, E.R. and McCauley, W.D. The effect of controIled three-dimensional roughness on
boundary layer transition at supersonic speeds, J. of the Aerospace Sciences, Vol. 27, N 4, 1960
69. Van Driest, E. R. and Blumer, C. B. Boundary layer transition at supersonic speeds: roughness effects
with heat transfer, AIAA Journal, Vol. 6, N 4, pp. 603-607, 1968
70. Owen, F.K. and Horstman, C.C. Hypersonic transitional boundary layers, AIAA Journal, Vol. 10, N
5,pp. 769-775,1972
71. Owen, F.K. Transition experiments on a flat plate at subsonic and supersonic speeds, AIAA Journal,
Vol. 8, W 3, pp. 518-523, 1970
72. Chen, K.K. and Thyson, N.A. Extension of Emmons' spot theory to flows on blunt bodies, AIAA
Journal, Vol. 9, W 5, pp. 821-825,1971
73. Morkovin, M.V. Transition to turbulence at high speeds, AFOSR Scientific Report, AFOSR-TR-701731, 1970
74. Amal, D., Coustols, E. and Juillen, J.C. Etude experimentale et theorique de la transition sur une aile
en fleche infinie, La Rechereche Aerospatiale 1984-4, 1984
75

Coleman, G.T., Elfstrom, G.M. and Stollery, J.L. Turbulent boundary layers at supersonic and
hypersonic speeds, AGARD Conference Proceedings N 93,1971

76

Poll, D.I.A. Transition description and prediction in three-dimensional flows, AGARD Report N 709,
1984

77

Amal, D. Three-dimensional boundary layers: laminar-turbulent transition, AGARD Report N 741,


1986

78

Saric, W.S. and Reed, HL Three-dimensional stability of boundary layers. Perspectives in turbulence
studies, Springer Verlag, 1987

79. Balakumar, P. and Reed, H.L. Three-dimensional stability of boundary layers, Phys. Fluids A3(4),
1991
80

Marcillat, J. Etude du developpement de la couche limite tridimensionnelle en regime de transition,


Technical Report W 19/2334, 1976

81

Stetson, K.F., Thompson, E.R., Donaldson, J.C. and Siler, L.G. Laminar boundary layer stability
experiments on a cone at Mach 8 - Part 3 : sharp cone at angle of attack, AIAA Paper N 85-0492,
1985

82

Creel Jr., T.R., Beckwith, I.E. and Chen, FJ. Transition on swept leading edges at Mach 3.5, J.
Aircraft, Vol. 24, W 10, 1987

264

Computational Methods in Hypersonic Aerodynamics

83

Amal, D. and Laburthe, F. Recent supersonic transition studies with emphasis on the swept cylinder
case, Conf. on Boundary Layer Transition and Control, Cambridge, 8-12 April, 1991

84

Malik, M. R. and Beckwith, I. E. Stability of a supersonic boundary layer along a swept leading edge,
AGARD CP N 438, 1988

85

Pfenninger, W. Flow phenomena at the leading edge of swept wings, AGARDograph 97, Part 4,1965

86

Poll, D. I. A. Boundary layer transition on the windward face of space shuttle during reentry, AIAA
Paper 85-0899,1985

87

Da Costa, J.L. Contribution a I'etude de la transition de bord d'attaque par contamination en


ecoulement hypersonique, Thesis, Poi tiers University, 1990

88

Morrisette, E.L. Roughness induced transition criteria for space shuttle-type vehicles, J. Spacecraft,
Vol. 13, W 2, 1976

Chapter 8:
Second-Order Effects in Hypersonic Laminar
Boundary Layers
J. Ph. Brazier, B. Aupoix, J. Cousteix
ONERA-CERT-DERAT, 2 avenue Edouard Belin,
B.P. 4025, 31055 Toulouse Cedex, France
ABSTRACT
In hypersonic flows, the bow shock wave in front of the body is the cause of a vortical inviscid
flow in the shock layer. The conventional boundary layer theory of Prandtl can not deal
with it. Van Dyke has derived an extended boundary layer theory using matched asymptotic
expansions where several second order effects as outer flow vorticity, displacement, or wall
curvature are accounted for. However, the matching between the boundary-layer flow and
the external inviscid flow is not perfect with second order expansions when the Reynolds
number is low and thus the boundary layer thick. To improve this and ensure a smooth
merging of the viscous flow into the inviscid one, a defect approach coupled with asymptotic
expansions has been proposed. Calculations on simple two-dimensional hypersonic bodies
are performed using the two methods and compared to full Navier-Stokes solutions.
INTRODUCTION
An hypersonic flow past a blunt body is characterized by a bow shock wave, detached in
front of the nose. For an inviscid flow, the entropy remains constant along a streamline,
except through the shock wave, where the entropy jump depends on the local slope of the
shock, and thus varies from one streamline to another one because of the shock curvature.
An entropy gradient orthogonal to the streamlines is thus created in the shock layer. On
the rear part of the body, the shock wave tends to be straight and the entropy gradient
disappears. The entropy gradient is thus confined in the vicinity of the wall in a region
called "entropy layer". This layer represents the streamlines which crossed the curved shock
wave near the nose.
This entropy gradient is related to the vorticity of the flow field through Crocco equation :
c~-;l iT 1\ iT = - grad H t + T grad S
This means that the streamwise component of the velocity varies through the entropy layer
along the normal to the body (fig. 1). Since the total enthalpy is constant in the whole flow,
the temperature varies too.
A reliable computational method for viscous flows must include the influence of the
vortical inviscid flow on the boundary layer. Solving the whole Navier-Stokes equations

266

Computational Methods in Hypersonic Aerodynamics

Figure 1: entropy layer


remains expensive, especially when real gas models are involved, and is not convenient for
design tasks. This is however necessary when the whole shock layer is viscous; but for
large enough Reynolds numbers, the boundary layer is sufficiently small compared to the
shock layer to assume that viscous effects are negligible near the shock wave. One can then
consider some distinct calculations for the inviscid shock layer and the viscous boundary
layer.
The classical boundary layer theory was first established by Prandtl, using an order-ofmagnitude analysis. According to this theory, the variables at the edge of the boundary layer
are matched with the inviscid variables at the wall. This is valid for large Reynolds numbers
and thus very thin boundary layers. One can then neglect the evolution of the inviscid
velocity and temperature throughout the boundary layer. Unfortunately, in hypersonic
flows, the Reynolds number is often moderate, because of the low density of the gas, and
the boundClIY layer thickness becomes no longer negligible compared to the entropy layer
thickness. The entropy layer may even be completely "swallowed" by the boundary layer
on the rear part of the body. A correct matching cannot then be obtained between the
boundary layer and the inviscid flow.
Van Dyke [18, 19, 20, 21] extended this boundary layer theory in 1962, using matched
asymptotic expansions. His higher-order boundary layer theory allowed him to bring into
evidence several second order effects. But when limited to second order expansions, this
theory cannot ensure a good matching at the edge of the boundary layer when the inviscid
profiles are not linear. To remedy this and ensure a smooth merging of the viscous flow into
the inviscid one, a defect approach coupled with matched asymptotic expansions has been
proposed. These two methods with examples of applications to two-dimensional flows will

Computational Methods in Hypersonic Aerodynamics 267


be presented here.
VAN DYKE'S THEORY
Navier-Stokes equations
We consider a steady flow of ideal gas past a convex blunt body. To derive boundary layer
equations, one must use curvilinear coordinates ((",'It), where {* is the distance along the
body and .,,* is normal to the body surface. All the variables are used in dimensionless form.
In the following, the starred variables are dimensional and the unstarred dimensionless. The
reference quantities are:
u~

free stream velocity

Ro

characteristic length, for example the nose radius

p~

free stream density

TO'

= U;'; ICp

fLo

= fL*(TO')

reference temperature
reference viscosity

and then the dimensionless variables are :


longitudinal coordinate: {=
normal coordinate : ."

~~

.,,*
= R*

longitudinal velocity: u

= ;:
00

normal velocity: v

v*

=-

U;"

p*

density: p = -

p~

pressure: p

p*

=U2
P* *
00

temperature: T

tY : fL

VISCOSI

enthalpy: h

00

= -T*
TO'

= -fL*
fLo
=

h*
U*2
00

These variables remain bounded in the stagnation region when the Mach number tends
towards iniinity (Van Dyke [19]). For an ideal gas, we then get:
h= T

268

Computational Methods in Hypersonic Aerodynamics

The Navier-Stokes equations for a two-dimensional flow write: (from [2])


- continuity:

- (-momentum:

- 7]-momentum :

- energy:
1 oh
pu hl o(

oh

op

op

+ pv 07] = hl o( + v 1}7]
1 I) [ I-'
h3 I}T]
1 I} [I-'
I}T]
+hlh31}( Pr Re hl I}( + hlh31}7] Pr Re hlh31}7]

I-'
+Re

[2 (hI1I}u
V I}hl)2
2(OV)2 2( u I}h 3 V I} h 3)2
I)( + hl 1}7]
+ 1}7] + hlh3 I}( + h3 1}7]
1 I}v

+ ( hI I)( +

(U ))2

I)

11}7]

hI

2 ( 1 I}u

3" hI

- state:

I}(

I}v

+ 1}7] + hlh3

,-I

p= - - p T

- sound velocity:

f! = j

(r - I)T

In these equations the following parameters appear :


- Prandtl number: Pr

= To

_ Reynolds number: Re

I-'*C*

= Poe* U*': R*0


1-'0

I}h3
I}(

V I}hl
1}7]

+ hl

V I}h3 )2]
1}7]

+ h3

Computational Methods in Hypersonic Aerodynamics 269


- metric coefficients :
hl

= 1 + ~ = 1 + K.TJ

where R( {) is the longitudinal curvature radius of the body, and r( {) is the transverse
curvature radius; a({) is the angle of the tangent to the body with the axis of symmetry
(fig. 2). For a plane flow j = 0 whereas for an axisymmetric flow j = 1.

r(FJ

_ _ _ _ -a....-_

l/K(r,;)

4-~

Figure 2: coordinates system


The boundary conditions are:
infinity:
u=l

v=Q

p=l

P = ,M;"

wall :

v=O

T= Tw

where Tw is the wall temperature. Instead of fixing the wall temperature, one can as well
impose the wall heat flux, or a relation between the temperature and the flux. The velocity
slip and temperature jump effects happen at the wall only for rarefied gases and will not be
accounted for here. Van Dyke [19) showed them to be second-order effects.
Asymptotic expansions
Two approximations of the Navier-Stokes solutions will be sought for: one far from the
wall, in the region where the viscous effects are weak, and another one near the wall, where

270

Computational Methods in Hypersonic Aerodynamics

the viscous effects are important. These approximations are looked for as expansions in
powers of a small perturbation parameter :
1

e:= ffe=

p.(To)
1
p.(T00) VReoo

Van Dyke [19J showed e: to be the right similarity parameter in the hypersonic limit when
both the Mach and Reynolds numbers tend towards infinity.
The outer expansions read:

u((,.,,)

= UI ((,.,,) + e:U2 ((,.,,) +

v((,.,,)

= VI ((,.,,) + e:V2((,.,,) +

p( ( , .,,)

= PI ( ( , .,,) + e: P2 ( ( , .,,) +
= R1 ((,.,,) + e: R2(e,.,,) +
= TI(e,.,,) + e:T2(e,.,,) +

p((,.,,)
T(e,.,,)

In the inner region, a stretched normal coordinate fi = .,,/e: is used according to the principle
of least degeneracy (Van Dyke [21, 19]). In this way, the normal coordinate is referred to a
quantity of the same order of magnitude as the boundary layer thickness, so that the new
normal coordinate is of order unity in the boundary layer. Similarly, a special expansion
must be written for the normal velocity because this quantity is of order e: in the boundary
layer, according to the continuity equation. Therefore, the inner expansions have the form:

u(c.,,)
v(e,.,,)

= Ul(e, fi) + e: U2(C fi) +


= e:i\(e, fi) + e: 2V2(e, fi) +

p(e,.,,) =Pl((,fi)

+ e:P2(e,fi) +

p((,.,,) = PI(e, fi)

+ e:P2(e, fi) +
+ c:i2(e, fi) +

T((,.,,)

= t1(e, fi)

All the coefficients and their derivatives are assumed to be of order unity.
Outer and inner expansions
These expansions are brought in the dimensionless Navier-Stokes equations and the like
powers of e: are equated. This leads to the following equations:

Computational Methods in Hypersonic Aerodynamics 271


- outer region, first order:

Here are the well-known Euler equations for an inviscid flow.


- outer region, second order:
continuity:

(-momentum:

7]-momentum :

energy:

state:

P2

,-1

= --(RIT2 + R2Tl)

These are just the Euler equations linearized for small perturbations. No viscous terms
appear until third order in the outer region.

272 Computational Methods in Hypersonic Aerodynamics


- inner region, first order:

These are the usual Prandtl equations. Note that the curvature radius only appears in the
continuity equation.
- inner region, second order:
continuity:

(-momentum:

TJ momentum:

energy:

state:

---------

Computational Methods in Hypersonic Aerodynamics 273


In these equations appear several source terms due to the longitudinal and transverse curvatures, and particularly a normal pressure gradient.
Matching conditions
The boundary conditions are not sufficient to solve both the inner and outer problems. The
missing conditions are obtained by matching the two expansions, i.e. writing that they are
equivalent in an intermediate region. Matching conditions are given by Van Dyke using his
matching principle :
The m-term inner expansion of the n-term outer expansion is equal to the n-term
outer expansion of the m-term inner expansion.
For the longitudinal component of the velocity, the two-term outer expansion is

Rewritten in internal coordinates

and expanded for e

-+

0, it yields:

This must be equivalent to the limit of the inner expansion for

r; -+

00 :

Equating the like powers of e leads to :

(1)
The second-order inner term is thus matched with the wall value of the second-order outer
term, plus the slope at the wall of the first-order outer term. The slope at the wall of the
second-order term arises only in third-order expansion, with the wall value of the secondorder derivative of the first-order term. The same argument is also valid with p, p and T.
The case of v is slightly different:

So:

or:

274 Computational Methods in Hypersonic Aerodynamics


This gives the boundary conditions at the wall for the external inviscid flow. The first one
is the classical slip-condition at the wall for an inviscid flow. The second represents the displacement of the second-order external flow by the first-order boundary layer. The normal
velocity of the inviscid flow is no longer zero, but has a fixed value, as if the wall was not
a solid line but a blowing was performed across it. This last condition is used for coupling
inner and outer calculations. So the first-order boundary layer must be solved before the
second-order external flow. The matched asymptotic expansions technique provides complete matching conditions between the inner and outer expansions, for all the variables at
any order, with a more rigorous argument than in Prandtl's theory. Figure 3 summarizes
the matching schemes for first- and second-order expansions. The matching between the
boundary layer and the external flow with second-order expansions is not perfect when the
external flow departs from a linear profile with constant slope.

first order

second order
Figure 3: matching schemes

The other boundary conditions are given by the free-stream conditions for the oncoming
flow, and the no-slip condition and fixed temperature at the wall :
at infinity:

Pa = 0

Computational Methods in Hypersonic Aerodynamics 275


at the wall (neglecting velocity slip and temperature jump) :

Due to these boundary conditions, the calculations must be chained as following:


outer flow, first order (Euler)
boundary layer, first order (Prandtl)
outer flow, second order (perturbed Euler)
boundary layer, second order
According to the equation :

OPl

or,

we get Pl(,7]) = Pl(,O). The pressure is thus constant across the boundary layer and
equal to the inviscid flow pressure at the wall.
A single composite expansion valid in the whole domain can be written as the sum of
the internal and external expansions minus the common part:

Thus the second-order composite expansion is :

This development is equivalent to each expansion in the corresponding region. The slope at
the wall of the velocity and temperature profiles is now equal to the boundary layer slope
plus the inviscid flow slope.
Second order effects
The second-order boundary layer equations are linear and thus the source terms and the
boundary conditions can be split in several parts corresponding to different second order
effects (Van Dyke [19]) :
longitudinal curvature (terms proportional to K)
transverse curvature (terms proportional to j cos Q)
displacement
entropy gradient
total enthalpy gradient
velocity slip
temperature jump

276

Computational Methods in Hypersonic Aerodynamics

These effects can be solved separately, except the displacement and entropy gradient effects
which are coupled, since in the second-order outer equations appears the term

which is the product of the displacement velocity by the shear of the first order inviscid
flow. Thus, for a given first-order boundary layer, the second-order flow will be different
depending on whether the external flow is sheared or not.
DEFECT APPROACH
Decomposition
As written above, the second-order matching is not perfect when the external inviscid profiles
are not linear across the boundary layer. This may happen when the Reynolds number is
moderate and thus the boundary layer is thick. To ensure a smooth matching at any
order whatever the external flow, a defect approach has been used, coupled with asymptotic
expansions. In the inner region, the variables are no longer the physical variables, but the
difference of them with the external solution (Le Balleur [13]). So we write:

= PE + PD

= VE + VD

- VE(('O)

= PE + PD
T = TE + TD
P

where the subscript E stands for the external variables and the defect variables are labelled
D. The term VE(CO) has been added to keep the condition VD(('O) = O.
Expansions are then written using the same small parameter e as Van Dyke, with the
coordinates ((,71) for the external functions and the stretched ((, r;) coordinates for the
defect variables:

= Ul((,71) + eU2((,71) +
VE((,71) = Vl(~,71) + eV2(~,71) +
PE((,71) = Pl((,71) + eP2((,71) +
PE((,71) = Rl(~,71) + eR2(~,71) +
TE((,71) = T1 ((,71) + eT2(~,71) +
UD(~,71) = Ul((,r;) + eU2(~,r;) +
VD((,71) = eih((,r;) + e2v2(~,r;) +
PD((,71) = Pl((,r;) + ep2((,r;) +
PD((,71) = Pl((,r;) + ep2((,r;) +
TD(~,71) = tl(~,r;) + et 2((,r;) +
UE(C71)

Computational Methods in Hypersonic Aerodynamics 277


Using a Taylor expansion, one can write near the wall :

VI(~,1/) - VI(~,O)

= eVi(~,1/)

V2(~,1/) - V2(~,O)

= eV2(~,1/)

The other coefficients are assumed to be of order unity, as well as their derivatives relative
to (~, 1/) for the external variables and (~, 7)) for the defect ones. This supposes that the
normal gradients in the inviscid flow are lower than those in the boundary layer, since they
do not have the same scale of magnitude.
Equations
In the outer region, the defect variables are null and the equations for the outer flow are
exactly the same as for Van Dyke's theory, i.e. Euler equations. Concerning the inner region,
one must first bring the above expansions into the Navier-Stokes equations, then substract
the external equations, and then equate same powers of e. For practical convenience, the
inner equations can then be rewritten in outer coordinates, replacing 7) by 1/, and with

Then the following first-order equations are obtained :


- continuity:

- {-momentum:

- 1/-momentum :

- energy:

- state:

PI

=' -

1 [PITl

+ (RI + pdtl]

As in Prandtl equations, the walJ curvature appears in the first-order equations only through
the transverse curvature radius in the continuity equation. The second-order equations are
small-perturbations of the above ones plus source terms due to curvature effects, like in

278

Computational Methods in Hypersonic Aerodynamics

Van Dyke's theory:


- continuity:

:~ [r j (P1U2 + P2U1) + rj(R1 + P1)U2 + r j (R2 + P2)U1]


+ :~ [rIJRej cos a(P1U1 + (R1 + P1)U1)]
() [.
.]
()V2 )
+ ()",
rJ P1 ( V2 - V2(~, 0) + V2 )
+ rJ P2(V1
+ vl) + rJ. ({)V1
R27h; + Rl7h;

+rj (V1 JRe -

+:'"
-

V2(~' 0)) (){)~1

+ R1 :'" [",(j cos a + Kr3) (V1 JRe -

V2(~' 0))]

[",JRe(jcosa+Krj)P1(V1+vd] =0

~-momentum

~
()~ + (R1 + pl)(U1 + U1) 8~
[(R1 + P1)(U2 + U2) + (R2 + P2)(U1 + ud ] ~

8U1
[
] 8U2
+ [P1U2 + P2 U1 + (R1 + Pl) U2 + (R2 + P2)U1 ] 8~ + P1U1 + (R1 + pdU1 8~

+ [R1(V1JRe -

V2(~'0)) + P1VRe(V1 + vd] (8~1

+ KU1)

+(R1 + pd(V1 + V1) :'" (U2 + K",JRe U1)


+ [(R1 + pd(V2 -

V2(~'0) + V2)

= -8P2
-+
8~

+ (R2 + P2)(V1 + V1)]

~~

8U1 ) + - 1 -() [K",vRe


r:;;8U1]
- 1 -8 ({)U2
/1-1- + /1-2/1-1Re 8",
8",
8",
Re 8",
8",

U
8/1-1 (8U1
+1- /1-1j cos a 8 1 + -1- JRe r
8",
JRe 8",- -8",- - ",-momentum:

(U1 + U1 ))

Computational Methods in Hypersonic Aerodynamics 279


- energy:
[(R I

+ PI)(U2 + U2) + (R2 + P2)(UI + uI) -

+ [p t U2
+(Rt

+ P2 UI + (RI + pI)U2 + (R2 + P2)UI -

+ PI)(UI + ud] ~~

K"7VRe (PIUI

+ (RI + PI)U I )] ~t

2 [
] 8T2
+ pI)(Ut + uI) 8t
8{ + PtUt + (Rt + pI)Ut 8{

+ [(RI + PI)(V2 +[Rt

It"7VRe(RI

V2({,O)

8t2
+ V2) + (R2 + P2)(Vt + VI) ] 8tt
8"7 + (Rt + PI)(VI + vd 8"7

(vl\/ae - V2 ((,O)) + PI (VI + vdVRe] : t


=Ut

8P2
8{

+ (U2 -

IT>'":) 8P
8P2
8
+(vtvRe
- V2({'O) 8"7 + (Vt + vda;;- + 8"7
8 [

+8"7

IJ.t ] 8Tt
Pr y'Re' 8"7

+1J.2

(8UI) 2 + 2IJ.t BUt 8U2


Re 8"7
Re 8"7 8"7

PE

--+

PE

--+

TE

8P2

+ (Ut + ud8[

( . cos

)-r-

8
+ 8"7

[ 1J.2 8tl]
Pr Re 8"7

and so for the defect variables :

UD

--+

VD

--+

VE({,O)

PD

--+

PD

--+

TD

--+

IJ.t

+ K + Pr y'Re'

2IJ.t BUt Bu t _ 21t1J.1 (UI

Matching conditions
At the edge of the boundary layer, we can write:

--+

8Pt

[ IJ.t 8t2]
Pr Re 8"7

VRe 8"7 8"7

- state:

. IT>'":

"7 lt v Reu t) 8{

y'Re'

8t t
8"7

+ ud But
8"7

280 Computational Methods in Hypersonic Aerodynamics


Thus at first order:

=0
lim Pl = 0
ij-+oo
lim tl = 0
1j-+00
lim Pl = 0
11-+00
lim Ul

11-+00

and at second order :


lim U2

ij-+oo

=0

=0
_lim P2 = 0
'1-+ 00
lim t2

tj-+oo

The wall conditions are:

hence:

Ul(~'O)
U2(~'O)
Vl(~'O)
tl(~'O)

t2(~'O)

= -Ul(~'O)
= -U2(~'O)
= V2(~,O) = 0
= Tw - Tl(~,O)
= -T2(~'O)

Discussion
Like in Van Dyke's theory, the coupling between the boundary layer and the external flow
is done by blowing at the wall. The chaining of the calculations is thus identical.
Using the above conditions, the first order 1/-momentum equation reduces to
Pl

=0

So, the pressure in the first-order boundary layer is everywhere equal to the local inviscid
flow pressure, instead of its wall value like in Van Dyke's theory. The second order defect
pressure is null too only on a plane surface, or neglecting curvature effects.
The defect boundary layer equations are parabolic and can be solved by space marching
at a very low cost, like the standard Prandtl equations.
From a theoretical point of view, it can be shown that the defect expansions are consistent with Van Dyke's ones by the fact that at a given order they differ only by terms which
are higher-order in Van Dyke's theory. If the flow over a plane surface is considered, the

Computational Methods in Hypersonic Aerodynamics 281


defect and outer first-order equations can be combined to form equations for the physical
variables in the inner region :

a(pu)

a(pv)

ar-+ {f;1
au
pu ae

+ pv a1J -

au

aUE
(pv - PEVE) 01J

aT
pu a(

+ pv a1J

aT

aTE
- (pv - PEVE) a."

a [

+a."

p, aTD]
Pr Re a."

p
which reduce to Prandtl's equations only when the inviscid flow is constant along the wall
normals. Note that only the defect variables appear in the viscous terms.
EXAMPLES OF APPLICATIONS
To experiment the defect approach, the first tries have been performed for an incompressible
shear flow past a flat plate. Second order calculations have been made with interesting results

([6, 7, 4, 5]).
Concerning compressible flows, several cases have been selected for a blunt body in a
hypersonic flow of ideal gas. The general shape of the body is a plane or axisymmetric
hyperboloid, defined by the nose radius and the angle of the asymptots, at zero degree incidence. The numerical data are given by Shinn, Moss and Simmonds [17] for an hyperboloid
equivalent to the windward symmetry line of the U.S. space shuttle. Two points of the
reentry trajectory of the STS-2 flight are considered here:
Reentry trajectory - Flight STS-2
26,6
Mach Moo
time (s)
250
altitude (km)
85,74
1,322
nose radius Ro (m)
asymptotes half-angle (0)
41,7
0,3634
pressure poo (Pa)
199
temperature Too (K)
7530
velocity Uoo (m/s)
6,3487.10- 6
density Poo (kg/m 3 )
reference tempera.ture To (K)
56321
Reynolds number Re
small parameter e

= PooUooRo
p,(To)

= Re- 1 / 2

Reynolds number Reoo

= PooUooRo

23,4
650
71,29
1,253
40,2
4,0165
205
6730
6,7979.10- 5
44900

183,55

1865,65

0,074

0,023

4792

42374

P,oo

The Prandtl number is assumed constant and equal to 0.725. The ratio of specific
heats I is 1.4. The wall temperature is fixed a.nd equal to 1500 K. The viscosity law is

282

Computational Methods in Hypersonic Aerodynamics

Sutherland's. No comparison with experimental data is possible since the real gas effects are
not yet included. So Navier-Stokes solutions [12] have been taken as reference, to compare
the two Euler + boundary layer methods. Euler calculations are made with a code from
ONERA [23]. Classical boundary layer solutions are obtained using a program developed in
DERAT ([3]). Only first-order boundary layer are presented here since second-order outer
:flow solutions are not yet available. Several second-order calculations using Van Dyke's
theory have been made on a hypersonic blunt body [1, 8, 9,10,11,14,15, 16].
Axisymmetric hyperboloid
In this case, the entropy layer is characterized by a non-zero entropy gradient at the wall and
a decreasing thickness towards the rear. Boundary layer profiles are displayed on figures 4
to 7 for the case Mach 23.4 . Longitudinal velocity are plotted on figure 4 at a distance of
four nose radius from the stagnation point. One can see on this figure the important velocity
gradient at the wall in the inviscid :flow. This gradient diminishes distinctly between the wall
and the boundary layer edge. So even with a second-order expansion, Van Dyke's method
could not give a good matching, since it considers only the wall value of the gradient. In
this case, it would widely overestimate the skin friction (Adams [1]). Due to the very low
wall temperature compared to the inviscid :flow one, the displacement effect is quasi-null
and the Navier-Stokes solution recasts exactly the inviscid profile in the outer region. In
this case, the agreement is excellent with the first-order defect boundary layer, better than
Van Dyke's composite expansion.
The corresponding profiles for the normal velocity and the temperature are shown on
figures 5 and 6. The very slight discrepancy observed between the Euler and Navier-Stokes
profiles confirms that the displacement effect is very low in this case. So the first order
defect approach gives very good results. Figure 7 shows the temperature profiles at nine
nose radius. The growing boundary layer has overlapped a larger part of the entropy layer.
The defect profile is in rather good agreement with the Navier-Stokes solution, but in this
case the composite expansion written with Van Dyke's first order solutions gives very bad
results and does not improve the inner solution. This is due to the negative slope at the wall
for the inviscid temperature. Because of the constant total enthalpy, the positive velocity
gradient at the wall induces a negative temperature gradient. In spite of this, the wall
heat :flux is increased by the vorticity as the skin friction, as can be seen on the figures 8
and 9. But the increase is far more important for the wall friction than for the flux. The
defect approach underestimates slightly these quantities but gives better predictions than
the classical boundary layer.
Figures 10 and 11 show the velocity and temperature profiles with an arbitrary temperature of ten times the temperature of the preceding case. The displacement effect is then far
more important and it is obvious on these figures that the N avier-Stokes solution is shifted
from the Euler solution in the outer zone. So the first-order boundary layer methods give
poor results and a second-order calculation seems to be necessary.
The velocity and temperature profiles at four nose radius abscissa in the Mach 26.6
case are presented on figures 12 and 13. Because of the lower density, the Reynolds number
is small and the boundary layer is about twice as thick as in the Mach 23.4 case. So a
large part of the entropy layer is overlapped by the boundary layer. The inviscid velocity
and temperature gradients at the edge of the boundary layer are far weaker than their wall
values. Due to the high value of the expansion parameter c, the second order effects are
more important and a slight displacement effect is visible between the Euler and Navier-

Computational Methods in Hypersonic Aerodynamics 283


Stokes profiles outside the boundary layer. The agreement between the Navier-Stokes and
defect profiles is rather good, but the shear at the wall is a bit too high for the later one.
Note that because of the negative inviscid temperature gradient at the wall, the Van Dyke's
composite expansion gives again poor results on the temperature profile.
The corresponding skin friction and wall heat flux are shown on figures 14 and 15.
As forecast from the velocity profiles, the defect approach improves greatly the standard
boundary layer result, but slightly overestimates the skin friction on the rear of the body.
In this region, due to the thinning of the entropy layer, the velocity gradient in the inviscid
flow and the boundary layer are of the same order of magnitude and the hypothesis done in
the defect theory does not stand any longer.
Plane hyperbola
Let us now consider a plane hyperbola in the same conditions of hypersonic flows. The
main difference with the axisymmetric case is that now the entropy gradient is null at the
wall, as showed Van Dyke [19J. SO the velocity and temperature gradients in the inviscid
flow are null at the wall too, and their influence will be significant only with a very thick
boundary layer. Moreover, far downstream, the flow can be assimilated to a parallel flow
and the entropy layer's thickness is constant whereas in the axisymmetric case the entropy
layer gets thinner towards the rear part of the body, since for a constant mass-flow the
circumference of the body increases. Thus the entropy gradient effects will now be far less
important.
Figures 16 and 17 show the velocity and temperature profiles on the hyperbola at nine
nose radius. The inviscid gradients are hardly visible outside the boundary layer and all the
methods give the same results.
When the Reynolds number is lower, the matching of the boundary layer with the
inviscid flow takes place in the gradient region, as shown on the figures 18 and 19 for the
case Mach 26.6. The defect method gives a good matching and a correct agreement with
Navier-Stokes solutions, but the two boundary layer methods give similar results near the
wall. Thus no significant difference is visible on the skin friction and the wall heat flux
shown on figures 20 and 21.
CONCLUSION
The different cases presented here showed on the one hand the validity of boundary layer
methods to compute hypersonic flows and on the other hand the importance of taking into
account the second order effects when calculating boundary layers at low Reynolds numbers.
The most important of them are the entropy gradient effect and the displacement effect.
They can deeply modify the wall quantities such as the skin friction or the wall heat flux,
which is essential to design the thermal protection of the vehicle.
Using the matched asymptotic expansions technique, the defect approach allows us to
improve the classical higher-order boundary layer theory of Van Dyke, for a similar cost.
Particularly it ensures a smooth matching of the viscous and inviscid flows, even when the
inviscid profiles vary significantly through the boundary layer. When the wall temperature
is low and thus the displacement effect is negligible, first-order defect calculations can give
good results and reproduce Navier-Stokes solutions with a reasonable accuracy at a lower
cost. It could then be a valuable tool for design tasks.

284 Computational Methods in Hypersonic Aerodynamics

Eulerhl ..der
0.12

Euler 18' order

Von Dyl<e hI O<d


composite ,,1 O<der

defect 1.1 order


\

0.10

defect hI order
0.10

Van Dyk.,.1 order


\
\

\
\

0.08
0.08

\
\

\
\

'" 0.08

I
/

0.02

. " ,,"

0.1

0.3

,
\

0.02

0.5

0.4

0.6

O.OO+----r----..,..------r---_+_
.(l.016

0.020

0.000

Van DyI<. ,.1 ...d..

0.18

defect 1.1 order

.(l.006

Eulerl.l ...d ..
oompoolte

compoolte 1.1 ..-der

0.010
v

Figure 5: Normal velocity profiles


Mach 23.4, Tw = 1500 K, { = 4

0.18

Von DyI<e 101 ...der

Eulerl.l ...der

0.10

'..

Figure 4: Longitudinal velocity profiles


Mach 23.4, Tw = 1500 K, { = 4

0.12

'.

0.04

""",,/

0.2

\
\

O.OO~:IU'><-'T---r--""'--"""'-'--"""--T"""

0.0

,,
,,
I

"'0.06

0.04

NavlerSloke.

NavlerStak..

1., ...

der

defect 1.1 order

NavI...Stak..

NavI ...Stak..

0.14

0.12

0.08

'" 0.08
I

0.04

/
/

./

,.,.''i!'''

0.02

"."

,/

,,
,,,

/./

0.1

0.2
T

0.08
0.06

/'

0.04

,."

0.02

o.oo+_.oo<"""'--r----.----.-----r'
0.0

",0.10

0.3

Figure 6: Temperature profiles


Mach 23.4, Tw = 1500 K, { = 4

0.4

0.00
0.0

0.1

0.2
T

0.3

Figure 7: Temperature profiles


Mach 23.4, Tw = 1500 K, { = 9

0.4

Computational Methods in Hypersonic Aerodynamics 285

0.007

Von Dyke 1st order


defect 1st order
~

0.006

Novier-Stokes

\ 0

0.005

~ 0.004

00
\

000
,

:E
~ 0.003

000

000

"

00000

' ........ ........

0.002

........ ........

---- ------

---- ----

----

0.001

0.0001+-----,------r-------.-----.-o
5
10
15
20
distonce oIong the body

= 1500 K

Figure 8: Skin friction on the hyperboloid - Mach 23.4, Tw


Von Dyke 1st order

0.006

defect 1st order


o

Novier-Stokes

0.005
0.004
)(

c2

0.003

----

0.002
0.001
0.000
0

10

15

20

distonce oIong the body

Figure 9: Wall heat flux on the hyperboloid - Mach 23.4, Tw

= 1500 K

286 Computational Methods in Hypersonic Aerodynamics

0.14

.....

Euler 1 st order

Van Dyke 1s1 order


defect 1st order

0.12

Navier-Stokes

0.10

0.08

""
,

,
,,
,,"""

0.04

0.02

"

Euler 151 order

I
I

Van Dyke 1s1 order

defect 1st order

0.12
0

I
I

Navler-Stokes

I
I

I
I

0.10

I
I
I

I
I
I
I
I

0.08

""

0.06

0.14

I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I

:.0

,
I

0.04

'0 :
. I

0/

,," "0

0.02

,-,,"0

"
0.00

0.00
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Figure 10: Longitudinal velocity profiles


Mach 23.4, Tw = 15000 K, = 4

0.4

I
I

Van Dyke 1st order

0.0

0.1

Figure 11: Temperature profiles


Mach 23.4, Tw = 15000 K, = 4

Euler 1st order

0
Van Dyke 151 order

composite 1st orde r

I
I

defect 1st order

I
I
I

Navier-Stokes

I
I

0.4

0.3

0.2
T

0.4

Euler 1st order

0.3

o
0.06

composite 151 order

defect 151 order

0.3

Navler-Stokes

I
I

I
I

""0.2

I
I

I
I

I
I

I
I

I
I
I

,
"
/'

,,'
"

,,

0.1

,.."

0.1

/.

"

,," """
.

O.O.;IlCI~r---'---.-----'~---r--...-0.0
0.1
0.2
0.3
0.4
0.5
0.6

Figure 12: Longitudinal velocity profiles


Mach 26.6, Tw = 1500 K, = 4

0.0+--a'IP<>W:"'--',.------.-----,----,--'0.0
0.1
0.3
0.4
0.2
T

Figure 13: Temperature profiles


Mach 26.6, Tw = 1500 K, = 4

Computational Methods in Hypersonic Aerodynamics 287

0.025

Van Dyke 1st order

,
\

0.020

defect 1st order


o

\
\

Navier-Stokes

\
\
\

:S.50!

0.015

.e

" 0.010

.ii

,
" , ...........

....... ......

0.005

-- ---- -----------

--------

O.OOO+----,r---,-----,---,-----,---,-----,---,-o
2
4
6
8
10
12
14
16
distance ciong the body

Figure 14: Skin friction on the hyperboloid - Mach 26.6, Tw

= 1500 K

o
Van Dyke 1st order
\ 0

0.020

defect 1st order

0
\

Novier-Stokes

0.015

,
"

00
0 0

....
'........

....

00
000

.............
......

0.010

...

000

--- --

-- --- ----------

0.005

O.OOO-+---r__-__,..--...--~--_r_--r__-__,..--..._-

6
8
10
12
distance along the body

14

Figure 15: Wall heat flux on the hyperboloid - Mach 26.6, Tw

16

= 1500 K

288

0.30

Computational Methods in Hypersonic Aerodynamics

.. ... Euler 11. order

0.30

. . . .. Euler I order
-_. Von Dyke 11. order

-_. Van Dyke 11' order


defect I order

0.26

defect I order

Navte,S\ok..

0.20

s:-

0.26

0.20

s:-

0.16

0.16

0.10

0.10

0.06

0.06

'T--,.--,.--..--..--._-._-'-

O.OO ..........
0.00 0.06

0.10

0.16

0.20
u

0.26

0.30

0.36

Figure 16: Longitudinal velocity profiles


Mach 23.4, Tw = 1500 K, ~ = 9

=-,.----.---.,-----,.....-:

0.00+...........
0.0

0.1

0.2

0.4

Figure 17: Temperature profiles


Mach 23.4, Tw = 1500 K, ~ = 9

Euler 11. order

.. ... Euler I order

Von Dyke I order

- _. Van Dyke 11. order

Navie,Stok..

0.3
T

defect I order

0.8

Navie,S\ok..

defect 11. order


o

0.8

Navte,Stok..

0.6

0.8

0.4

0.4

0.2

0.2

0.0-+,1oLI:"-.,..---,---._-C--.,---0.2
0.3
0.0
0.1
0.4
u

o.o++';loO.<~.-----r---._--_"T--'

Figure 18: Longitudinal velocity profiles


Mach 26.6, Tw = 1500 K, ~ = 9

0.0

0.1

0.3

0.2
T

Figure 19: Temperature profiles


Mach 26.6, Tw = 1500 K, ~ = 9

0.4

Computational Methods in Hypersonic Aerodynamics 289

Von Dyke 1st order


defect 1st order

0.010

Novier-Stokes

0.008

.~

:g 0.006
.10

0.004

0.002

O.OOO~---------r---------r---------r---------r

10

15

20

distonce along the body

Figure 20: Skin friction on the hyperbola - Mach 26.6, Tw


0.014

= 1500 K

Von Dyke 1st order


defect 1st order

0.012

Novier-Stokes

10

15

0.010

)(

0.008

oi!
0.006
0.004
0.002
0.000
0

20

distance along the body

Figure 21: Wall heat flux on the hyperbola - Mach 26.6, T",

= 1500 K

290 Computational Methods in Hypersonic Aerodynamics

References
[lJ Adams J .C. Higher-order boundary-layer effects on analytic bodies of revolution, Arnold
Engineering Development Center, Report AEDC-TR"68-57 (1968)
[2J Anderson D.A., Tannehill J.C., Pletcher R.H. Computational fluid mechanics and heat
transfer, Mc Graw-Hill (1984)
[3J Aupoix B., Arnal D. CLIC: calcul des couches limites compressibles, Rapport technique
OA 23/5005 AYD (1988)
[4J Aupoix B., Brazier J.Ph., Cousteix J., Monnoyer F. Second-order effects in hypersonic
boundary layers, Third Joint Europe-U.S. Short Course on Hypersonics, Aachen (FRG),
October 1-5 (1990)
[5J Aupoix B., Brazier J .Ph., Cousteix J. An asymptotic defect boundary layer theory applied to hypersonic flows AIAA 91-0026, 29th Aerospace Sciences Meeting, January 710, Reno, Nevada (1991)
[6J Brazier J .Ph., Aupoix B., Cousteix J. Etude asymptotique de La couche limite en formulation dejicitaire, Compte-Rendus de l'Academie des Sciences, t. 310, serie II, pp.15831588 (1990)
[7J Brazier J .Ph. Etude asymptotique des equations de couche limite en formulation
dejicitaire, PhD thesis, Ecole Nationale Superieure de l'Aeronautique et de l'Espace,
Toulouse, France (June 25 1990)
[8J Davis R.T., Fliigge-Lotz I. The laminar compressible boundary-layer in the stagnationpoint region of an axisymmetric blunt body including the second-order effect of vorticity
interaction, International Journal of Heat and Mass Transfer, Vol. 7, pp. 341-370 (1964)
[9J Davis R.T., Fliigge-Lotz I. Laminar compressible flow past axisymmetric blunt bodies,
Journal of Fluid Mechanics, Vol. 20, Part 4, pp. 593-623 (1964)
[10J Fannel6p T.K., Fliigge-Lotz I. Two-dimensional hypersonic stagnation flow at low
Reynolds numbers, Z. Flugwiss. Vol. 13, nO 8, pp. 282-296 (1965)
[l1J Fannel6p T.K., Fliigge-Lotz I. Viscous hypersonic flow over simple blunt bodies; comparison of a second-order theory with experimental results, Journal de Mecaruque, Vol.
5, nO 1, pp. 69-100 (1966)
[12J Lafon A. Calcul d'ecoulements visqueux hypersoniques, Rapport technique DERAT
nO 32/5005.22 (1990)
[13J Le Balleur J .C. Calcul des ecoulements a forte interaction visqueuse au moyen de
methodes de couplage, AGARD - CP 291 (1980)
[14J Marchand E.O., Lewis C.H., Davis R.T. Second-order boundary-layer effects on a slender blunt cone at hypersonic conditions, AIAA 6th Aerospace Sciences Meeting, AIAA
Paper nO 68-54 (1968)
[15J Maslen S.H. Second-order effects in laminar boundary layers, AIAA Journal Vol. 1,
nO 1, pp. 33-40 (1963)

Computational Methods in Hypersonic Aerodynamics 291


[16] Papenfuss H.D The second-order boundary-layer effects for the compressible threedimensional stagnation-point flow, Journal de Mecanique, Vol. 16, nO 5, pp. 705-732
(1977)
[17] Shinn J.L., Moss J.N., Simmonds A.L. Viscous shock layer heating analysis for the
shuttle windward plane with surface finite catalytic recombinations rates, AIAA Paper
n 82-0842 (1982)
[18] Van Dyke M. Higher approximations in boundary layer theory, Part 1: General analysis,
Journal of Fluid Mechanics, vol. 14, pp. 161-177 (1962), Part 2: Application to leading
edges, Journal of Fluid Mechanics, vol. 14, pp. 481-495 (1962)
[19] Van Dyke M. Second-order compressible boundary-layer theory with application to blunt
bodies in hypersonic flow, Hypersonic Flow Research, Vol. 7, pp. 37-76 (1962), F.R.
Riddell Editor, Academic Press
[20] Van Dyke M. Higher-Order boundary-layer theory, Annual Review of Fluid Mechanics
(1969)
[21J Van Dyke M. Perturbation methods in fluid mechanics, Parabolic Press (1975)
[22] Vasantha R., Nath G. Semi-similar solutions of the insteady compressible second-order
boundary-layer flow at the stagnation point, International Journal of Heat and Mass
Transfer, Vol. 32, nO 3, pp. 435-444 (1989)
[23] Veuillot J.P. Calcul de l'lkoulement axisymetrique autour d'un corps de rentree, Rapport Technique ONERA nO RT 31/1285 AY (Janvier 1989)

Chapter 9:
Real Gas Effects in Two- and Three-Dimensional
Hypersonic, Laminar Boundary Layers
B. Aupoix, J. Cousteix
ONERA/CERT, Aerothermodynamics Department,
2 avenue E. Belin, B.P. 4025, 31055 Toulouse Cedex,
France
ABSTRACT
This paper is devoted to hypersonic laminar boundary layers as the maximum heat
loads during the re-entry of a space shuttle occur in the laminar regime. The emphasis is placed upon the real gas effects which deeply affect the evolution of the
flow. As boundary layers are addressed, thermal non-equilibrium effects are neglected
and the attention is focused upon chemical non-equilibrium effects. A simplified
atom/molecule model is justified and more complex models are detailed.
The influences of the wall catalysis, of the scatter in chemical reaction rate data
and of the wall temperature are brought into evidence.
Self similar solutions, to determine the flow and the wall heat flux at two- and
three-dimensional stagnation points or along attachment lines are presented.
INTRODUCTION
The design of a space shuttle heavily relies upon CFD. One of the key problems is
the evaluation of the heat loads the vehicle will experience. An Euler plus boundary
layer approach is well suited to give such information for a reasonable computational
cost. This paper is devoted to the problem of real gas effects which occur in hypersonic flows as the hypervelocity flow upstream of the shock is transformed into
a hyperenthalpy flow in which dissociation and ionization phenomena occur. It is
restricted to laminar boundary layers as the maximum heat loads are encountered in
the laminar regime. The entropy layer swallowing phenomenon, which must be accounted for to correctly predict the heat loads, is addressed in a companion paper [14].
INTRODUCTION TO REAL GAS EFFECTS IN RE-ENTRY FLOWS
The source of real gas effects
The gas stagnation enthalpy is conserved through a shock wave. For hypersonic flows,
a high velocity, low temperature flow upstream of the shock wave is transformed into a

~------------------------------------------------------------------------------------~!-

296 Computational Methods in Hypersonic Aerodynamics

hI
h

enthalpy of a species I per mass unit


enthalpy of the mixture per mass unit

We have the following relationships:


(1)

(2)

p=nM

Thermodynamics of a gas mixture


For boundary layer applications, we will only
consider gases in thermal equilibrium. The enthalpy of a species I reads

hI =

h~ + loT Cp[dT

(3)

where C p [ is the specific heat of the species I at constant pressure per mass unit and
h~ the heat of formation of the species I which, by convention, is zero for molecular
oxygen and nitrogen.
For a gas mixture, assuming an ideal mixture, the mixture enthalpy is the sum of
the contribution of each species, i.e. as h is the enthalpy per mass unit

(4)
The specific heat of the mixture should read

(5)
I.e. from (4)

(6)
where the first term of the RHS is the frozen specific heat

CPf = LCICp [

(7)

i.e. the specific heat of a mixture of constant composition while the second term
usually does not make sense since the gas composition does not depend only upon the
gas temperature and pressure, except for gases at chemical equilibrium.
All energy levels of particles are quantified. It is therefore possible to describe how
energy is stored through an energy partition function which includes the contribution
of all energetic modes. Thermodynamic functions such as the internal energy, the
enthalpy or the entropy can then be related to this partition function. As the energy
levels can be deduced from spectroscopic measurements, they are known with great
accuracy.

Computational Methods in Hypersonic Aerodynamics 297

State equation for a gas mixture


gas, i.e.
R

Each species is assumed to behave as a perfect

R = 8.314 Jmol- 1 K- 1

PI = PI MI T

(8)

and the Dalton's law yields


P= LPI
I

= p-T

(9)

If Mo is the molar mass of air upstream of the shock i.e. before any chemical reactions,
the state equation can be expressed with the help of the compressibility factor Z as
P = ZpRoT

Z= Mo
M

(10)

A brief introduction to air chemistry


Gas phase chemistry
Collisions between particles lead to chemical reactions. If
colliding particles have a sufficient relative translational energy, the two particles may
get close enough during the collision process to form an activated complex which may
give new chemical species. What appears as a reaction at the macroscopic level is in
fact a set of elementary steps at the microscopic level, with first a collision, formation
of activated complex, parting of new species and perhaps third body collision. This
description may itself be a simplification of the complete process, as a reaction may
in fact be a chain reaction formed of elementary reactions. All collisions cannot lead
to a reaction as there exists an energy barrier to overcome. If the relative translation
energy of the two particles is too weak, they cannot form an activated complex.
Species can be required for the reaction to occur but are not modified at the end
of the reaction process, as the third body in the atom recombination process or the
collision partner in the dissociation reactions. These species can however control the
effectiveness of the reaction process and act as a catalyst.
For dissociation problems, it is common practice to only take into account molecular and atomic nitrogen and oxygen and nitrogen monoxide (N2' O2, NO, N, 0)
and to neglect argon, carbon dioxide and water vapour.
The relevant reactions for studies of flows around bodies during re-entry are [33,
37,38]
Dissociation reactions
N2 +M ;=: N+N+M
(11 )
O2 +M ;=: O+O+M
NO+M ;=: N+O+M
where M stands for any of the species .
Shuffle reactions
NO+O ;=: N+0 2
(12)
o +N2 ;=: N+NO
It must be noticed that the direct formation reaction for the nitrogen monoxide

N 2 +0 2 ;=:NO+NO

(13)

introduced in early models (see e.g. [36]) is now usually discarded as it can be obtained
as a combination of dissociation and shuffle reactions. This gives a system of seventeen
chemical reactions.

298 Computational Methods in Hypersonic Aerodynamics


For ionization studies, the nitrogen monoxide gives the major part of the electrons for re-entry flows up to 6000ms- 1 . Simple models only consider ionized nitrogen
monoxide and the electron (NO+, e-). For velocities larger than 9000ms-I, the ionization is mainly due to atomic nitrogen and oxygen. A general ionization chemical model
takes eleven species into account, i.e. N 2, O 2 , NO, N, 0, N;, ot, NO+, N+, 0+, e-)
The ionized species can now act as third bodies in the dissociation reactions (11) and
new chemical reactions are to be accounted for [33, 37, 38]
Charge exchange reactions

o +ot
N 2+N+
o +NO+
N 2+O+
N+NO+
02+ NO+
N+NO+
o +NO+

--->.
.,--

O 2 +0+

--->.

N +N;
NO+O+
0 +N;
NO+N+
NO+Ot
0 +N;
O2 +N+

.,--

--->.
.,---->.
.,--

;=:
--->.
.,---->.

.,---->.

.,--

(14)

Associative ionization reactions

N+N

;=:
;=:
;=:

0+0

N+O

N; +eot +eNO+ +e-

(15)

O++e-+eN+ +e- +e-

(16)

Electron impact ionization reactions

O+eN +eChemical reaction rates

;=:
;=:

Let us consider a simple reaction:

A+ B

--+

AB

The rate of production of the species AB is proportional to the number of possible


collisions between reacting particles. By noting WAB, WA, WB the variation per time
unit of PAB, PA, PB due to chemical reactions, we have:

(17)
where WAB/MAB represents the variation per time unit of nAB due to chemical reactions. Since, from the atom conservation nA + nAB = constant and nB + nAB =
constant, we also have:

Arrhenius [2] proposed to express the temperature dependence of the reaction rate
constant by his celebrated law

(18)

Chapter 9:
Real Gas Effects in Two- and Three-Dimensional
Hypersonic, Laminar Boundary Layers
B. Aupoix, J. Cousteix
ONERA/CERT, Aerothermodynamics Department,
2 avenue E. Belin, B.P. 4025, 31055 Toulouse Cedex,
France
ABSTRACT
This paper is devoted to hypersonic laminar boundary layers as the maximum heat
loads during the re-entry of a space shuttle occur in the laminar regime. The emphasis is placed upon the real gas effects which deeply affect the evolution of the
flow. As boundary layers are addressed, thermal non-equilibrium effects are neglected
and the attention is focused upon chemical non-equilibrium effects. A simplified
atom/molecule model is justified and more complex models are detailed.
The influences of the wall catalysis, of the scatter in chemical reaction rate data
and of the wall temperature are brought into evidence.
Self similar solutions, to determine the flow and the wall heat flux at two- and
three-dimensional stagnation points or along attachment lines are presented.
INTRODUCTION
The design of a space shuttle heavily relies upon CFD. One of the key problems is
the evaluation of the heat loads the vehicle will experience. An Euler plus boundary
layer approach is well suited to give such information for a reasonable computational
cost. This paper is devoted to the problem of real gas effects which occur in hypersonic flows as the hypervelocity flow upstream of the shock is transformed into
a hyperenthalpy flow in which dissociation and ionization phenomena occur. It is
restricted to laminar boundary layers as the maximum heat loads are encountered in
the laminar regime. The entropy layer swallowing phenomenon, which must be accounted for to correctly predict the heat loads, is addressed in a companion paper [14].
INTRODUCTION TO REAL GAS EFFECTS IN RE-ENTRY FLOWS
The source of real gas effects
The gas stagnation enthalpy is conserved through a shock wave. For hypersonic flows,
a high velocity, low temperature flow upstream of the shock wave is transformed into a

296 Computational Methods in Hypersonic Aerodynamics

hI
h

enthalpy of a species I per mass unit


enthalpy of the mixture per mass unit

We have the following relationships:


(1)

(2)

p=nM

Thermodynamics of a gas mixture


For boundary layer applications, we will only
consider gases in thermal equilibrium. The enthalpy of a species I reads

hI =

h~ + loT Gp]dT

(3)

where Gp] is the specific heat of the species I at constant pressure per mass unit and
h~ the heat of formation of the species I which, by convention, is zero for molecular
oxygen and nitrogen.
For a gas mixture, assuming an ideal mixture, the mixture enthalpy is the sum of
the contribution of each species, i.e. as h is the enthalpy per mass unit

(4)
The specific heat of the mixture should read

(5)
I.e. from (4)

"""

Gp = L..J GI
I

aaT
hI"",, aGI
+ L..J hI aT
I

"""
I

= L..J GIGp ]

aGI
+ """
L..J hI aT

(6)

where the first term of the RHS is the frozen specific heat
GpJ =

2: GIGp ]

(7)

i.e. the specific heat of a mixture of constant composition while the second term
usually does not make sense since the gas composition does not depend only upon the
gas temperature and pressure, except for gases at chemical equilibrium.
All energy levels of particles are quantified. It is therefore possible to describe how
energy is stored through an energy partition function which includes the contribution
of all energetic modes. Thermodynamic functions such as the internal energy, the
enthalpy or the entropy can then be related to this partition function. As the energy
levels can be deduced from spectroscopic measurements, they are known with great
accuracy.

Chapter 9:
Real Gas Effects in Two- and Three-Dimensional
Hypersonic, Laminar Boundary Layers
B. Aupoix, J. Cousteix
ONERA/CERT, Aerothermodynamics Department,
2 avenue E. Belin, B.P. 4025, 31055 Toulouse Cedex,
France
ABSTRACT
This paper is devoted to hypersonic laminar boundary layers as the maximum heat
loads during the re-entry of a space shuttle occur in the laminar regime. The emphasis is placed upon the real gas effects which deeply affect the evolution of the
flow. As boundary layers are addressed, thermal non-equilibrium effects are neglected
and the attention is focused upon chemical non-equilibrium effects. A simplified
atom/molecule model is justified and more complex models are detailed.
The influences of the wall catalysis, of the scatter in chemical reaction rate data
and of the wall temperature are brought into evidence.
Self similar solutions, to determine the flow and the wall heat flux at two- and
three-dimensional stagnation points or along attachment lines are presented.
INTRODUCTION
The design of a space shuttle heavily relies upon CFD. One of the key problems is
the evaluation of the heat loads the vehicle will experience. An Euler plus boundary
layer approach is well suited to give such information for a reasonable computational
cost. This paper is devoted to the problem of real gas effects which occur in hypersonic flows as the hypervelocity flow upstream of the shock is transformed into
a hyperenthalpy flow in which dissociation and ionization phenomena occur. It is
restricted to laminar boundary layers as the maximum heat loads are encountered in
the laminar regime. The entropy layer swallowing phenomenon, which must be accounted for to correctly predict the heat loads, is addressed in a companion paper [14].
INTRODUCTION TO REAL GAS EFFECTS IN RE-ENTRY FLOWS
The source of real gas effects
The gas stagnation enthalpy is conserved through a shock wave. For hypersonic flows,
a high velocity, low temperature flow upstream of the shock wave is transformed into a

294 Computational Methods in Hypersonic Aerodynamics


low velocity, high enthalpy flow downstream. For a shuttle re-entry, at 70 km altitude
and Mach number about 25, the Rankine-Hugoniot formulae yield temperatures in
the shock layer about 25000 K. Of course, air is no longer an ideal gas at such temperature, real gas effects are encountered. Real gas effects will be briefly presented in
this section. The reader is referred to [1, 3, 15, 17, 20, 31, 39, 49J for more detailed
information.
Energy storage in air molecules
At room temperature, air is mainly made of molecular nitrogen and oxygen. These
molecules can store energy in various ways. The first mode is the translation kinetic
energy due to the Brownian motion of the particle. When the molecules are excited,
energy can be supplied to the electrons, so that they can move to higher energy levels.
At high excitation levels, the electron leaves the molecule, it is the ionization process.
Molecules can also rotate around their center of mass. At last, excited molecules can
vibrate, i.e. the link between two atoms allows them to move along the axis with
respect to each other. When the vibration energy is large, the atoms may part, it is
the dissociation process. The internal energy of the nucleus is neglected.
Atoms only have translation and electronic excitation energy levels.
Collisions between particles tend to homogenize the probability of the energetic
state of particles so that they tend towards thermal equilibrium in which energy is
"equally" distributed among particles and energetic modes.
At room temperature, energy is stored only on the translation and rotation modes.
Through the shock wave, the mean flow kinetic energy is transferred to the translation
mode. It requires a few collisions, i.e. about the thickness of the shock wave, for the
translation, rotation and electronic excitation modes to equilibrate. It may require
hundreds or thousands of collisions to equilibrate the vibration mode. So, behind
the shock wave, vibrational energy relaxation is observed, together with dissociation
reactions as the vibrational energy increases and even ionization if the energy level is
high enough.
Equilibrium and non-equilibrium flows
The so-formed gas mixture tends towards a new equilibrium state, from both a thermal and a chemical point of view. Thermal equilibrium is achieved when the energy is
distributed among all the energy levels according to the statistical mechanics, i.e. following Boltzmann distribution. Chemical equilibrium is achieved when, for constant
pressure and energy, the mixture has no tendency to undergo any change in chemical composition, no matter how slow. Both the thermal and the chemical relaxation
processes can be characterized by relaxation time scales which can be compared with
the evolution time scale of the flow motion.
Various situations occur according to the relative orders of magnitude of these
time scales .
When the relaxation time scale is very long compared to the motion time scale,
the flow is frozen. The relaxation does not proceed during the flow evolution,
the chemical composition or the vibrational energy remains constant .
When the relaxation time scale is very short compared to the motion time scale,

- - -

~-----

---

Computational Methods in Hypersonic Aerodynamics 295

the flow is in equilibrium. The chemical composition or the vibrational energy


immediately adapts itself to changes in the local conditions .
When the relaxation time scale and the motion time scale are of the same order
of magnitude, the flow is a non-equilibrium flow.
Thermal and chemical relaxations, while occuring together and competing downstream of a shock wave, can have very different time scales. A flow can be, e.g.
at equilibrium from a thermal point of view while frozen from a chemical point of
VIew.

Thermal relaxation
Boundary layers will be addressed in the following. For the boundary layer concept
to be valid, an inviscid region must exist between the shock wave and the outer edge
of the boundary layer. For re-entry flows, vibrational relaxation occurs downstream
of the shock wave and is achieved at the outer edge of the boundary layer. The flow
in the boundary layer can be assumed at thermal equilibrium.
It must however be pointed out that this thermal equilibrium assumption neglects
two physical phenomena:
When atoms recombine at the wall, the so-formed molecules can be out of vibrational equilibrium. As this vibrational non-equilibrium quickly relaxes in the
wall region, this phenomenon can be neglected .
Free electrons due to ionization reactions hardly interact with atoms and molecules so that their translation energy is out of equilibrium with the atom and
molecule translation energy.
In the following, we shall restrict ourselves to flows at thermal equilibrium, which
are relevant to boundary layer flows.
Thermodynamics of a gas mixture
Characteristics of a gas mixture Before considering the thermodynamic properties
of a gas mixture of species (noted by capital indices I), let us define some quantities
we will need later:

V:
mI:

m
PI:
P

CI

nI:

n :
MI :
M
PI
P

volume occupied by the mixture


mass of a species I contained in V
mass of the mixture contained in V
density of a species I : PI = mdV
density of the mixture: P = mjV
mass fraction of a species I : C I = PI j P
number of moles of a species I per volume unit
number of moles of mixture per volume unit
molar mass of a species I
molar mass of the mixture
partial pressure of a species I
pressure of the mixture

296 Computational Methods in Hypersonic Aerodynamics

hI
h

enthalpy of a species I per mass unit


enthalpy of the mixture per mass unit

We have the following relationships:


p=

LPI

(1)

(2)

p=nM

Thermodynamics of a gas mixture


For boundary layer applications, we will only
consider gases in thermal equilibrium. The enthalpy of a species I reads

hI =

h~ + loT Cp[dT

(3)

where Cp [ is the specific heat of the species I at constant pressure per mass unit and
h~ the heat of formation of the species I which, by convention, is zero for molecular
oxygen and nitrogen.
For a gas mixture, assuming an ideal mixture, the mixture enthalpy is the sum of
the contribution of each species, i.e. as h is the enthalpy per mass unit

(4)
The specific heat of the mixture should read

(5)
I.e. from (4)

Cp

" hI aC
" hI aC
= '"
L.J CI ah
aT + 'L.J
aT = '"
L.J C[C p[ + 'L.J
aT
[

(6)

where the first term of the RHS is the frozen specific heat

CPJ = LC[Cp[

(7)

i.e. the specific heat of a mixture of constant composition while the second term
usually does not make sense since the gas composition does not depend only upon the
gas temperature and pressure, except for gases at chemical equilibrium.
All energy levels of particles are quantified. It is therefore possible to describe how
energy is stored through an energy partition function which includes the contribution
of all energetic modes. Thermodynamic functions such as the internal energy, the
enthalpy or the entropy can then be related to this partition function. As the energy
levels can be deduced from spectroscopic measurements, they are known with great
accuracy.

Computational Methods in Hypersonic Aerodynamics 297


State equation for a gas mixture
gas, I.e.
PI

Each species is assumed to behave as a perfect

= PI MIT

(8)

and the Dalton's law yields


P

LPI

(9)

p-T

If Mo is the molar mass of air upstream of the shock i.e. before any chemical reactions,
the state equation can be expressed with the help of the compressibility factor Z as
Z= Mo
M

P = ZpRoT

(10)

A brief introduction to air chemistry


Gas phase chemistry
Collisions between particles lead to chemical reactions. If
colliding particles have a sufficient relative translational energy, the two particles may
get close enough during the collision process to form an activated complex which may
give new chemical species. What appears as a reaction at the macroscopic level is in
fact a set of elementary steps at the microscopic level, with first a collision, formation
of activated complex, parting of new species and perhaps third body collision. This
description may itself be a simplification of the complete process, as a reaction may
in fact be a chain reaction formed of elementary reactions. All collisions cannot lead
to a reaction as there exists an energy barrier to overcome. If the relative translation
energy of the two particles is too weak, they cannot form an activated complex.
Species can be required for the reaction to occur but are not modified at the end
of the reaction process, as the third body in the atom recombination process or the
collision partner in the dissociation reactions. These species can however control the
effectiveness of the reaction process and act as a catalyst.
For dissociation problems, it is common practice to only take into account molecular and atomic nitrogen and oxygen and nitrogen monoxide (Nz, Oz, NO, N, 0)
and to neglect argon, carbon dioxide and water vapour.
The relevant reactions for studies of flows around bodies during re-entry are [33,
37,38]
Dissociation reactions

N z +M
Oz +M
NO+M
where M stands for any of the species .
Shuffle reactions

NO+O
+Nz

~
~

~
~

N+N+M
O+O+M
N+O+M

(11 )

N+O z
N+NO

(12)

It must be noticed that the direct formation reaction for the nitrogen monoxide
(13)

introduced in early models (see e.g. [36]) is now usually discarded as it can be obtained
as a combination of dissociation and shuffle reactions. This gives a system of seventeen
chemical reactions.

298 Computational Methods in Hypersonic Aerodynamics


For ionization studies, the nitrogen monoxide gives the major part of the electrons for re-entry flows up to 6000ms- l . Simple models only consider ionized nitrogen
monoxide and the electron (NO+, e-). For velocities larger than 9000ms- l , the ionization is mainly due to atomic nitrogen and oxygen. A general ionization chemical model
takes eleven species into account, i.e. N 2, O2, NO, N, 0, Ni, ot, NO+, N+, 0+, e-)
The ionized species can now act as third bodies in the dissociation reactions (11) and
new chemical reactions are to be accounted for [33, 37, 38]
Charge exchange reactions

o +ot
N 2+N+
o +NO+
N 2+O+
N+NO+
02+NO+
N+NO+
o +NO+

---"

"<""""""

---"
"<""""""

---"
"<""""""

---"
"<""""""

---"
"<""""""

---"
"<""""""

---"
"<""""""

---"
"<""""""

O2 +0+
N +Ni
NO+O+
0 +Ni
NO+N+
NO+Ot
0 +Ni
O2 +N+

(14)

Associative ionization reactions

N+N ~ Ni +e0+0 ~ ot +eN+O ~ NO++e-

(15)

Electron impact ionization reactions

O+eN +eChemical reaction rates

~
~

O++e-+eN+ +e- +e-

(16)

Let us consider a simple reaction:

A+B

-t

AB

The rate of production of the species AB is proportional to the number of possible


collisions between reacting particles. By noting WAB, WA, WB the variation per time
unit of PAB, PA, PB due to chemical reactions, we have:

(17)
where wABI MAB represents the variation per time unit of nAB due to chemical reactions. Since, from the atom conservation nA + nAB = constant and nB + nAB =
constant, we also have:

=
Arrhenius [2] proposed to express the temperature dependence of the reaction rate
constant by his celebrated law

(18)

Computational Methods in Hypersonic Aerodynamics 299

where EA is the activation energy. This form has later been extended, to cover a
wider range of reactions, as

= ATE exp ( - ~; )

(19)

which is sometimes referred to as non-Arrhenius behaviour.


Up to now, only irreversible chemical reactions have been considered. However,
chemical reactions encountered in re-entry flows may proceed in both ways.
Using the Penner notation, a chemical reaction can be written as

where f and b indicate the forward and backward reactions, the A are the chemical
species and the v the stoechiometric coefficients. The difference VI = v'f - VI must
be introduced as it indicates whether the species AI is destroyed or produced by the
forward reaction. The evolution of the reactants is given by the following expressions
which are generalisations of equation (17)
Forward reaction

(20)
Backward reaction

,,"
nI=-kbvI II (nJ)J

(21)

where k j and kb are respectively the forward and backward reaction constants. The
chemical production rate of species I is the balance between the two terms

(22)
Non-Arrhenius expressions for the reaction rate constants can be found in various
publications, either from reviews of data for combustion [7, 24, 51] or from papers
dealing with re-entry problems [12, 16, 28, 29, 33, 36, 37, 38, 50]
The reaction rate constants are mainly determined by experimental techniques.
It must be pointed out that such experiments are very difficult to design and analyze.
Reaction rate data are thus known with little accuracy, especially at high temperatures. Various sets of reaction .rate constants, namely one used by Straub [47] and
due to Oertel [36], one due to Park [40, 37], one used by Shinn et al. and due to
Bortner [12], the one recommended by the National Bureau of Standard [51] i.e. the
one due to Baulch [7] and, at last, the review due to R. H. Hanson and S. Salimian
in Gardiner's textbook [24] are compared on figures 1 and 2. Figure 1 shows a very
large scatter for the oxygen dissociation reaction which is one of the most important
reactions. Even if Straub data are discarded, there is one order of magnitude discrepancy between Baulch and Park models. The ratio of reaction rate constants to the
value recommended by the National Bureau of Standard is plotted on figure 2 for a

300 Computational Methods in Hypersonic Aerodynamics


_10 '0

,..--Straub

<1'

-'

_ Park
_Sninn
Baulch

("0

~105

10-;~LO~0----------~2~0~0~O----~3~00~O~~LO~0~O~5~0~O~0--~~~8~O~OO~~
TI K I

Figure 1: Comparison of reaction rate constant determinations for the O 2


reaction

+ N2

-->

+ 0 + N2

J .0

,,
,,

'"
CD

'"o
"'2.5

..................
Straub

2.0
_ -

- Park

0.5

O.

1000

2000

3000

LOOO

5000

5000

1000

8000
TI K)

Figure 2: Comparison of reaction rate constant determinations for the NO


reaction

+0

-->

+ O2

shuffle

Computational Methods in Hypersonic Aerodynamics 301

shuffle reaction. The dotted lines indicate the evaluated uncertainty range [51]. The
scatter is larger than estimated, a ratio about three between extreme values can be
observed.
Further information about the forward and backward reaction rate constants can
be obtained by considering chemical equilibrium. Chemical equilibrium is obtained
for a reaction when the forward and backward reactions exactly balance. As the
production is zero, the reaction rate constants can be linked to the number of moles
of species per volume unit
"I
kf
(23)
nI eq
= k

II(
I

or, using the state equation PI = nIkT where k is the Boltzmann constant,

II (PI eq)"I = kkf (kT)L.I"I = J<p (T)


I

(24)

which is known as the mass action law. This law can also be deduced from thermodynamical arguments, as the equilibrium state corresponds to the minimum of the
Gibbs free energy of the system formed by the species involved in the chemical reaction. The equilibrium constant J<p and the mixture composition can be determined
from thermodynamical data. The ratio of the forward and backward reaction rate
constants can be obtained so; therefore it is accurately known from thermodynamical
data.
Chemical time scale According to the above formulae, the chemical time scale for
e.g. the forward reaction in equation (20) reads:
Tchem

rv

nI (kf

PI (kf

rv

(25)

-1

PI = nIMI :

or, introducing the density of the species I


Tchem

gn~~)
gP~~)

(26)

-1

We define the reaction order as the sum of the stoechiometric coefficients of the
reactants, i.e. for the forward reaction
I

= ~vJ

(27)

The chemical time scale reads:


Tchem

rv

C(k P'"
I

-1

gC;~)

-1

(28)

i.e. the chemical time scale decreases when the chemical reaction rate or the density
increases.
Let us consider a re-entry trajectory such as the STS-2 flight re-entry trajectory
the laminar regime part of which is given in table 1.

302 Computational Methods in Hypersonic Aerodynamics

t
(8)
200
250
330
460
480
540
650
770
830
1000
1120
1215

Altitude
(km)
92.35
85.74
77.91
74.98
74.62
73.33
71.29
68.67
66.81
60.56
52.97
47.67

u
(km 8- 1)
7.50
7.53
7.42
7.20
7.16
7.03
6.73
6.31
6.05
4.99
3.87
2.96

p
(kg m- 3 )
2.18410 6
6.36510- 6
2.33510- 5
3.18510- 5
4.05510- 5
4.79410- 5
6.82410- 5
9.66910- 5
1.21610- 4
2.62110- 4
6.76210- 4
1.34410- 3

p
(atm)
1.128 10 6
3.58710- 6
1.316 10- 5
2.14210- 5
2.28010- 5
2.83110- 5
3.96510- 5
5.99210- 5
7.92510- 5
1.87710- 4
5.02510- 4
9.90010- 4

(I<)
324
199
199
198
198
200
205
219
230
253
262
260

Mach
number
27.90
26.60
26.30
25.50
25.40
24.80
23.40
21.30
19.90
15.70
11.90
9.15

Angle of
attack
40.4
41.0
40.2
40.0
40.3
40.4
39.4
38.5
41.4
42.0
38.3
34.8

Table 1: STS-2 flight - re-entry trajectory

(from [46])

As the shuttle goes deeper into the atmosphere, the density increases by roughly
three orders of magnitude. As the order of the reactions is usually two or three, this
implies a variation of three or six orders of magnitude of the pv' -1 term. At the
same time, the shuttle slows down so that the flow enthalpy and the temperature
level decrease. In the outer region of the boundary layer, near the stagnation point,
the temperature decreases from about 6000 K to about 3000 K. The variations of
reaction rate coefficients with temperature can be very steep. For example, for the
atomic oxygen dissociation reaction, the forward reaction rate coefficient decreases by
four order of magnitude between 6000 K and 3000 K while the backward reaction rate
coefficient hardly changes. So, the chemical time scale strongly depends upon both
the considered chemical reaction via the temperature dependence of the reaction rate
coefficient and the altitude via the density.
For boundary layer flows, the flow is highly dissociated in the stagnation point
region and recombines downstream as the flow acceleration leads to a decrease in the
temperature level. As the recombination reactions are third order reactions, the variation of the density is the leading term and the chemical time scale decreases with
the altitude. On the other hand, the mean motion time scale, i.e. the length of the
shuttle divided by the upstream velocity, varies weakly. This results in an important
variation of the chemical relaxation/mean motion time scale ratio during the re-entry.
The flow is frozen from a chemical point of view at the beginning of the re-entry
(high altitudes, low densities) and near equilibrium at the end of the laminar regime
(lower altitudes, higher densities). We shall focus our attention here on chemical nonequilibrium flows, which occur over quite all the laminar part of the shuttle re-entry.
Gas/surface interface chemistry The surface material may act as a catalyst. The
heterogeneous catalysis phenomenon occurs in three basic steps.

Computational Methods in Hypersonic Aerodynamics 303


Atoms which diffuse towards the surface may be adsorbed at the surface. This
first step depends upon both the impinging atom and the surface: this is the
surface selectivity. For example, metal surfaces can easily adsorb oxygen atoms
while metal oxide surfaces are more relunctant to adsorb oxygen.
Once the atom is inserted inside a site of the surface, it may react with either
another atom adsorbed in another site (the Langmuir-Hinshelwood process) or
by collision with a free atom in the gas (the Eley-Rideal process). The wall
atoms the adsorbed atom is bound with act as third bodies as they provide or
dissipate the necessary or excess energy for the reaction. Sites are very efficient
third bodies as surface atoms are bound to other atoms.
The third step is the desorption of the new species from the wall, the surface site
is now free and the process can occur again.
Goulard [27J has proposed to model the wall catalysis with the help of an Arrhenius
law. The flux of species at the wall due to catalysis is expressed as
(29)

where the subscript w indicates wall values and the reaction order m should be between
one and two. Scott [44J assumes the reaction order to be equal to one, the reaction
rate constant klw has thus the dimension of a velocity.
As pointed out previously, the heterogeneous catalysis first depends upon the
ability of the surface to adsorb gas species. These various behaviours are reflected by
the values of the reaction rates coefficients k1w . When the coefficients tend towards
zero, the species are not modified at the wall. The wall is thus said to be non-catalytic.
On the contrary, when the coefficients tend towards infinity, the atoms impinging the
wall recombine very quickly. The wall is said to be catalytic. One may also assume
that the wall has a catalytic effect on all possible chemical reactions so that chemical
equilibrium is achieved at the wall. As the wall temperature is "low" (i.e. up to 1500 K)
chemical equilibrium means almost complete recombination of atoms at the wall as
air hardly dissociates even at low pressure for these temperatures. Therefore, the
atom dissociation energy is transferred at the wall; wall catalytic efficiency increases
the wall heat flux. The non-catalytic and catalytic wall cases are extreme cases which
are not encountered practically. Real materials have a finite catalytic efficiency. Here
again, experimental determination of the wall catalytic efficient is very difficult [13,
22, 43, 44J and moreover the wall catalytic efficiency evolves with the aging of the
surface material.
Concerning ionized species, it is taken for granted that they are neutralized at the
wall, whatever the wall catalytic efficiency.
Transport properties of gas mixtures
Collisions are also responsible for momentum exchange between particles which leads
to transport of momentum or heat by the gas mixture. They also tend to homogeneize the mixture composition via a diffusion process. Gas kinetic theory yields the

304 Computational Methods in Hypersonic Aerodynamics


expression for the shear stress

~=

~diV11.~) + 1]div11.~

fi (grad11. + grad t 11. -

(30)

for the mass diffusion flux


9.1

"

~P~

MIMJ D d
DTgradT
IJ-J- I -T
J#

(31 )

QI

nI+ (nI
PI) gradp
gra d --n
n
P
p

(32)

and for the heat flux

"

"" nJ

Dr (qI=- - =qJ)

1
q = -AgradT+ ~hN +RT ~~--I

-t

-I

n MI DIJ

PI

(33)

pJ

where fi is the viscosity, 1] the bulk viscosity, DIJ the binary diffusion coefficient of
species I in species J, D IJ the polynary diffusion coefficient of species I in species J in
the thermal diffusion coefficient and the thermal
presence of the other species,
conductivity.
The thermal diffusion term in the mass diffusion flux (31) corresponds to the Soret
effect i.e. the fact that light species diffuse towards hot regions, heavy species towards
cold regions. The thermal diffusion term in the heat flux (33) corresponds to the
Dufour effect. The thermal diffusion term is a second-order term in gas kinetic theory
so that these effects are small.
The various transport coefficients of the mixture can be expressed in terms of
collision integrals which are known with little accuracy at high temperatures.
Fick's law provides an approximate expression for the mass diffusion flux as

Dr

9.1 = - pDgradCI

(34)

It is also interesting to introduce dimensionless numbers which measure the ability


of the gas mixture to transport one quantity with respect to another. The usual
numbers are

Prandtl number
Lewis number
Schmidt number

p - fiCf/ _momentum diffusion


-

conduction
mass diffusion
X conductio.D
S - P - L_momentum dlttusion
- C - pD mass diffusion

.c -

pD PI _

(35)

Equilibrium air properties


In the absence of any constraint, gas mixtures tend towards thermal and chemical
equilibrium. It is then important to know the equilibrium state as it is the asymptotic solution. The following results for dissociated air are taken from [21], from
temperatures ranging from 1000 to 7000 K and pressures from 0.001 to 10 atmospheres.
Composition of air at chemical equilibrium Mass fractions of the main species are
plotted versus temperature and pressure on figures 3 to 7. Oxygen dissociates at

Computational Methods in Hypersonic Aerodynamics 305

temperatures of about 2500 K while nitrogen dissociates at higher temperatures,


about 5000 K, so that the oxygen dissociation is nearly complete when nitrogen dissociation starts to proceed. The lower the pressure, the more easily molecules dissociate.
Only small amounts of nitrogen monoxide can be obtained. The higher the pressure, the higher the nitrogen monoxide maximum mass fraction.
Thermodynamical properties of equilibrium air The evolutions of the reduced enthalpy, the frozen specific heat and the compressibility factor are plotted on figures 8
to 10. The two separate molecule dissociation processes are very clear on the evolution
of these quantities. When the compressibility factor is equal to two, all molecules are
dissociated.
Transport properties of equilibrium air The viscosity, thermal conductivity, Prandtl
and Lewis numbers are plotted on figures 11 to 14. As the temperature increases, the
viscosity deviates from the Sutherland law. As the pressure decreases, the dissociation is more important and both the viscosity and the thermal conductivity increase.
The Prandtl number, i.e. the ratio between momentum and heat transport, is weakly
affected by oxygen dissociation but varies with nitrogen dissociation. However, in the
considered range of temperatures, it can be assumed constant. The Lewis number
decreases with enthalpy or temperature, i.e. the commonly assumed value 1.4 is valid
at low temperatures but not for re-entry flows.
As the Prandtl and Lewis numbers are close to unity, the gas transports momentum, heat and mass in a similar way. Therefore, the dynamic, thermal and chemical
boundary layers have similar thicknesses.
GOVERNING EQUATIONS FOR REAL GAS FLOWS
The equations which govern the mixture flow are the state equation (9), the species,
continuity, momentum and energy equations. The expressions for the species diffusion, momentum and heat fluxes have been given above (30 to 33).
Species equations
Species equations express the balance between advection, chemical production and
species diffusion. They read
DC[
. - d'lvq
(36)
p
- - =WI
1)t

where

-I

J5t stands for the substantial derivative


1)

-1)t = -at +u.grad


-

(37)

These equations are linked as

(38)

306 Computational Methods in Hypersonic Aerodynamics

0.8-.-----.------,------r-----,r-----,------,

0.4-+------~----~-----+----~r-_+--_h~--~

0.2~----~~----~-----+------r4r_~rl_--~_i

1000

2000

3000

4000

5000

6000

7000

T (K)

Figure 3: Molecular nitrogen mass fraction at chemical equilibrium

0.8-.------r-----,------.------.-----,------,

0.6-+------~----~r_----_r------+_~--~----~~

0.4-+------~----~r_----_r------+_--+--,-+----~

0.0~~~~~~~~~~~~~~FT~~4-~~~

1000

2000

3000

4000

5000

6000
T (K)

Figure 4: Atomic nitrogen mass fraction at chemical equilibrium

7000

Computational Methods in Hypersonic Aerodynamics 307

0.25-r------.------,-------r------r-----~------~

0.20~-----4~~~~----+-----~----_+----~

0.15_r------r_-+~~~~--r_-----r------~----~
p
(atm)

0.10-+------+---4-~~~~~------+_----_4------~

0.05-+------r----+-#--~~r_~r__r------~----~

0.00-+~rT~~~rT~~~~~~~~~~~~~~
1000
2000
3eee
4000
5000
60e0
7000
T (K)

Figure 5: Molecular oxygen mass fraction at chemical equilibrium


0. 25-.------.------,r------.------.------r------,

0.20-+------+-----++~--~~~~~~~--~----~

p (atm)

0.15-+------+---~~~~_+~--+_~------~----~

e.10-+------+---+-~4_~~~~----+_----_4------~

0.05-+------+-4-+-++~~~+-----_r------r_----~

1000

2000

3000

,
4000

5000

60130
T (K)

Figure 6: Atomic oxygen mass fraction at chemical equilibrium

7000

308 Computational Methods in Hypersonic Aerodynamics

0.10-.------.-----~r-----_r------._----_,------~

0.08-+------+-----~~--~~~

__--+_----_4------~

0.06-+------+-----~~~--~------~----_4------~

0.04_+------~----~~----~~--_+----~~----~

0.02-+------+-~=_~~----~------~~--_4----~~

1000

2000

3000

4000

5000

6000

7000

T (Kl

Figure 7: Nitrogen monoxide mass fraction at chemical equilibrium


25~------~----~-------r------~----~------~

ROT

5-+------+-~~Lb~~~+-----~------+_----~

1000

2000

3000

4000

5000

6000

7000

Figure 8: Evolution of dimensionless enthalpy H/noT at chemical equilibrium

Computational Methods in Hypersonic Aerodynamics 309

100-.------.------,-------r------r-----~----~

80-+------~------+-------~----_4------~----~

60-+-------+-------+-------+----~-+~----~------~

1000

2000

3000

4000

5000

6000

7000
T (K)

Figure 9: Evolution of the specific heat coefficient Cp/1lo at chemical equilibrium


2.0-.----~r_----.-----_r------r_----~==p_~
Mo
Z=-

1.8-+------+------4------~------+-+_--~----~~

1.6-+------+------4------~------~~~_4~----~

1.4-+------+-----~------~--~--~--~~--~--~

1.0-*~~~~~~~~~~~~~~~~~~~~

1000

2000

3000

4000

5000

6000

7000

Figure 10: Evolution of the compressibility factor at chemical equilibrium

310 Computational Methods in Hypersonic Aerodynamics

e.25_.------.-----_.------.------r------r-----,
10 3 ~ (pl)

-,

0.20-+------4-----~~----~------+_-----1~~~~

e.15-+------4-------r------+------~_7~~~~--~

--- -e.10-+------4-------r-~~~--~~~~--~-------~

0.05~~~--~----~------~----~------~----~

0.00~,_"_r+,rT._~_.,_ro-r"rT._+o_.,_r+_r,,rT~

1000

2000

30013

4131313

6000

70013

Figure 11: Evolution of the viscosity at chemical equilibrium


0.5-r----~r-----,-----_r------r_----,-----~

e.4-+------~----1-----_+------t_~~~~-r~

e.3-+------+-----_+------+------YL;~~~~~~

0.2~------~----~----~~~---r----~r_----~

0.1~--~~~----+-----_+----~r_----1_----_1

10013

20013

30013

4131313

50013

E000

7000

Figure 12: Evolution of the thermal conductivity at chemical equilibrium

Computational Methods in Hypersonic Aerodynamics 311


0. 90-r------,------,r------,------,------.------,

0.85-+-------r------+-------r-----~------_r~nr-,y

~ Cpf

= --),-

0.80-+------+-----_+------+-~~_+--~~:r__r---i

0.75~----_;-------;-------;------~7__r--r+~~--~

1000

2000

3000

4000

5000

6eoo
7000
T (Kl
Figure 13: Evolution of the Prandtl number at chemical equilibrium
1. 6

Ii

JI{

lA

Cl

I~
o

1.2

()

Y-l
Ll

G9~
c

1.0

<)

p/P re !
.8

.0 I -

.4

0
0

10- 4
10-:1

10- 2

6
LI

10 0

1~

~
u

10- 1

10 1
Correlating !unctlon,
lable I

.2

.4

.0

.0

1.0

EnU,aJpy rallo,

1.2

1.4

K:n--!__

1.0

h/hE

Figure 14: Evolution of the Lewis number at chemical equilibrium

(Pref

= 1 atm, hE = 19.7106 J Kg-I)

(from [18])

loB

-(],

o iii

2. o

312 Computational Methods in Hypersonic Aerodynamics


Continuity equation

-dp
+ d'IVpU- =
dt

(39)

Momentum equation

p~~ = div (-p, +~)

(40)

Energy equation
Gas radiation effects will be assumed to be negligible. The energy equation can be
written in various forms, as well for the internal energy e as

e
p DD t = -divqt
-

+ (r.- -

s..

(41 )

P..)
- :-

as for the enthalpy

Dh

p Dt = -dIVlJ.t

Dp
Dt

+ <I>D

(42)

where <I> D is the dissipation function


(43)

Boundary layer equations for two-dimensional flows


For two-dimensional, plane or axisymmetric flows, the boundary layer equations read,
with x along the body, y along the normal, U and v the velocity components in these
two directions:
State equation

p=p-T
M

(44)

ap
1 apuRi
apv _ 0
at + Ri a;- + ay -

(45)

Continuity equation

with j = 0 in plane flow and j = 1 in axisymmetric flow. R is the distance of a point


of the wall to the body axis in axisymmetric flow.
Species equations
( 46)

(47)
Momentum equations

au
au
au
p- +pu- +pvat
ax
ay

ap aT

--+ax ay

ap
ay

au
ay

= 11-

( 48)

(49)

Computational Methods in Hypersonic Aerodynamics 313


Energy equation for the enthalpy
oh
oh
oh
oqty
op
op
(OU)
P-+PU-+PV-=--+-+U-+J.l &t
ox
oy
oy
ot
OX
oy
oT
qty = -,Xoy

'"' nJ- 1
Dr
+ '"'
L..Jq[yhI + RT '"'
L..JL..J
-I

Or, using the stagnation enthalpy H


oH
P ot

oH

J n M I 1)1J

(qlY
-- -qJy)
PI
PJ

(50)

(51 )

= h + ~ , the energy equation reads:


oH

+ pu Ox + pv oy

op
= ot

+ oy (UT -

qty)

(52)

Boundary layer equations for three-dimensional flows


For three-dimensional flows, the governing equations (36, 41) hold but the expression
of the advection term is somewhat more complex and the momentum equation must
be written for the three space directions. When these equations are reduced to their
boundary layer form, the boundary layer equations are similar to (44, 50) with now
two momentum equations along directions parallel to the wall, a more complex form
of the advection term in the species, momentum and energy equations and a modification of the source term in the energy equation. The final form of the equations
strongly depends upon the choice of a surface coordinate system. Readers are referred
to [30] for a discussion of surface coordinate systems and to [5, 11] for an example of
three-dimensional boundary layer equation set.
Boundary conditions
The boundary layer equations form a set of parabolic equations which can be integrated with a x-marching procedure once initial and boundary conditions have been
prescribed.
For two-dimensional flows, boundary layer profiles at the initial station and boundary conditions at the wall and at the outer edge of the boundary layer are required.
For three-dimensional flows, it must be reminded that any information diffuses
along the normal to the wall and is advected downstream by all the streamlines
crossing this normal. This means that a given point influences a large downstream
domain. Therefore, boundary layer profiles should be prescribed everywhere flow
enters the computation domain.
The boundary conditions at the wall reflect the wall imposed constraints. If the
velocity slip is neglected, the velocity is null at the wall. If the temperature jump is
neglected, the fluid temperature can either be prescribed equal to the imposed wall
temperature, or determined from an imposed wall heat flux. The heat flux a fluxmeter
measures is the sum of the heat flux qtyw due to the boundary layer flow and other
sources or sinks, mainly radiation effects. If the gas radiation is neglected and only
the wall radiation is accounted for, the measured flux reads:

(53)

314 Computational Methods in Hypersonic Aerodynamics


where c is the wall emissivity factor and a the Stephan-Boltzmann constant (a =
5.671O- 8 Wm- 2 K- 4 ). It has been observed during a long period in the space shuttle
re-entry that the wall temperature remains constant, i.e. if heat conduction inside
the wall is neglected, this flux is null. This is known as radiative equilibrium when
the radiated flux balances the boundary layer flux. When radiative equilibrium is
imposed as boundary condition, both the wall temperature and the wall heat flux are
results of the computation.
The wall boundary conditions for the species equations reflect the interactions
between the gas and the surface material which can act as a catalyst and promote
chemical reactions at the interface. The species diffusion flux at the wall must balance
the flux due to chemical reactions at the interface:
(54)
where the flux q[cat is given by (29) for atoms. Extra relations can be deduced from
the fact that interface reactions do not create or destroy chemical elements so that
the flux of chemical elements is null at the wall. For air dissociation problems, only
five species (N2' O 2 , NO, N, 0) are taken into account. This yields two element conservation relations for nitrogen and oxygen. Therefore, only three interface reaction
rates are independent. It is usually assumed that nitrogen monoxide is not modified
at the interface, and only the interface reaction rates for atomic nitrogen and oxygen
are specified. For a catalytic wall, it is possible to impose either complete atom recombination with no modification of the nitrogen monoxide or chemical equilibrium
at the wall. This is discussed at length in [45].
According to the matching asymptotic expansion approach [48], the flow outside
of the boundary layer must match with the inviscid flow at the wall. Outside of the
boundary layer, when all gradients along the wall normal vanish, the momentum (48),
energy (50) and species (46) equations reduce to their inviscid form at the wall. The
outer pressure, velocity, temperature and species distributions could be imposed as
the solution at the wall in an inviscid flow calculation. Practically, due to round-off
and integration errors, it is not possible to have a complete consistency between all
these prescribed external values. It is enough to impose either the velocity or the pressure field and to compute the missing external values by integrating the momentum,
energy and species equations with vertical advection and diffusion effects omitted.
The atom/molecule approximation
As pointed out previously, oxygen dissociation is almost complete when nitrogen dissociation starts to proceed. Moreover, only small amounts of nitrogen monoxide are
observed. The gas is roughly a mixture of molecular nitrogen, atomic and molecular oxygen at "low" temperatures; of atomic oxygen, atomic and molecular nitrogen
at "high" temperatures. Moreover, oxygen and nitrogen have similar molar masses,
transport and thermodynamical properties. The problem can be simplified by considering a mixture of only two species: atoms A and molecules A 2
Basic relations

The following relationships can easily be verified

Masses of the species MA2 = 2MA

Computational Methods in Hypersonic Aerodynamics 315


Mass fractions C A2 = 1 - C A so that most of the following variables will be
expressed in terms of CA
The partial pressures are expressed as

(55)

so that the stagnation pressure reads


where Z

= l+CA

The enthalpy per mass unit of each species reads

and since CPA


h =
=

rv

h~ + loT CPAdT

(57)

loT C PA2 dT

(58)

C PA2 , the mixture enthalpy reads

+ C A2 hA2 = hA2 + C A (hA - hAJ


+ CAh~ + C A faT (CPA - C PA2 )dT

(59)

CAhA
hA2

rv

hA2

+ CAh~

(60)

and the frozen specific heat is


(61)

Chemistry The chemistry of the atom/molecule mixture can be represented with a


single chemical reaction
(62)
A 2 +M;::=; A+A+M
where the third body M stands for both the molecule A2 and the atom A. The
chemical reaction rate can be obtained as

(63)
so that, introducing the chemical equilibrium values (subscript E), and with the auxiliary relations nM = nA + nA2' nA = pCA/MA the above equation reads
.

nA

p3
) C~ - C~E
= -kRM3
(1 + CA
C2
A
1AE

(64)

Transport
For a binary mixture, the mass diffusion can be represented with the
help of a Fick law (34) provided thermal diffusion effects are negligible and pressure
gradients are small. The transport properties of the mixture are thus usually represented by assuming constant Prandtl and Lewis numbers. The Sutherland law is

316 Computational Methods in Hypersonic Aerodynamics


sometimes used to compute the viscosity, as it leads to about 10% error for chemical
equilibrium air [23].
The various fluxes thus read
QA

I.

llt

C
pD = ptL

-pDgradYA

p~ -

tL (gradu. + gradtu. -

(65)

~divu.~)

= -; (gradh + (C -1) (hA -

- 77divl!

(66)

h A2 )gradCA )

with hA - hA2 '"

h~ (67)

It must be pointed out that this model is however a very crude one which requires
some tuning to give good results according to the considered range of temperature.
While CPN2 '" CP02 , the formation enthalpies are different as hfJv2 '" 2h~2 and the
reaction rate coefficients for oxygen and nitrogen dissociations are different.

TWO-DIMENSIONAL BOUNDARY LAYER FLOWS


Stagnation point solutions for atom/molecule approach
Fay and Riddell solution As heat transfer is often maximum at the stagnation point,
axisymmetric stagnation point solutions were investigated by Fay and Riddell [23]
using an atom/molecule approach. The boundary layer equations were transformed
with the help of the Levy-Lees-Dorodnitsyn space coordinate transformation [20]
77 =

RUe
V2f,

l
0

pdy

(68)

where the subscript w denotes wall values, U e is the velocity outside the boundary
layer, R the distance to the symmetry axis (R '" x near the stagnation point), x
and y the boundary layer coordinates respectively along the body and normal to the
wall. With this set of space variables, the boundary layer equations are written for
the reduced velocity, enthalpy and energy profiles
H
s=He

(69)

where H is the stagnation enthalpy and the subscript e denotes values outside of the
boundary layer.
Self similar solutions can be obtained at the stagnation point, the boundary layer
equations reduce to a set of coupled ordinary differential equations.
A reaction rate kb '" T1.5 was used and both frozen and equilibrium flows were
also investigated.
All the results presented herein are for the same conditions, i.e. a Lewis number of 1.4, a Prandtl number of 0.71, a wall temperature of 300 K and a wall enthalpy/external enthalpy ratio of 0.0123.
At the stagnation point, only the velocity component normal to the wall is not
null, the species mass fraction profiles thus result from a balance between advection
towards the wall, chemistry and diffusion. The atom mass fraction profiles for a frozen
flow on a catalytic wall are given on figure 15. For a frozen flow on a non-catalytic

---~----

Computational Methods in Hypersonic Aerodynamics 317


to
TE"rERA"iURE

09

~OUllI~RIUMr
FR~ZEN ~

OB

07

01

I/

o~
o 0.2

7'

1/ v
V

.........-: l7'
0.4

/ATOM MASS FRACTION C.

y< ~~~UILlB~IUM
t AOZ5N
1.0

OB

./

...-e:::P-:::--

0.6

I'" v/

03

02

0.6

05

.--

Ano 9

12

1.4

1.6

1.8

2.0

22

2.4

2.6

'1

Figure 15: Temperature and atom mass fraction profiles for frozen and equilibrium flows (from [23])

wall, the atom mass fraction is constant throughout the boundary layer and equal to
the external value. The wall catalytic efficiency modifies the boundary condition at
the wall and strongly affects the whole atom mass fraction profile.
The wall heat flux is also affected by the wall catalysis. From a study over a wide
range of altitudes from 7.5 km to 36 km and velocities from 1.8 to 7 kms-I, and for
different values of the Prandtl and Lewis numbers, Fay and Riddell proposed fits for
the wall heat flux as

cJ>w
hw -He
cJ>w
hw -He
cJ>w
hw - He
with

A=

A(1 + [co.
A(1 + [C

52 -

O 63
.

1] ~ee)

-1]

A(l-~:)
0.76

(p Pw/-tw
Pe/-te )

0.4

~:)

Chemical equilibrium flow

(7'))

Frozen flow - Catalytic wall

(71)

Frozen flow - Non-catalytic wall

(72)

(73)

(dUe)
P Pw/-tw dx 0

where the subscript 0 indicates values at the stagnation point and hD is the dissociation enthalpy per mass unit, i.e. the atom formation enthalpy times their mass
fraction.
These formulae bring into evidence the influence of wall catalysis upon heat flux.
Atoms diffuse towards the wall and, for a catalytic wall, recombine and release their
dissociation energy. This leads to an higher wall heat flux for the catalytic wall. As
the wall heat flux can be expressed as

8T
8CA
cJ>w = -.\ 8y - pD (hA - hA2 ) 8y

(74)

the flux on a non-catalytic wall is only due to the temperature gradient while both

318 Computational Methods in Hypersonic Aerodynamics


terms intervene for a catalytic wall. The contribution of both terms is given on
figure 16 on which the frozen flow corresponds to a zero recombination rate parameter.
The larger the recombination rate parameter, the more atoms recombine in the
boundary layer as shown on figure 17. On a non-catalytic wall, the more atoms recombine in the boundary layer, the more energy is released and the larger the wall
heat flux. On a catalytic wall, the more atoms recombine in the boundary layer, the
less energy is released at the wall due to catalytic reactions and hence the smaller the
wall heat flux as shown on figure 16.
Goulard extension to finite catalysis The study of frozen flow axisymmetric stagnation point was extended by Goulard [27] to account for finite catalytic recombination
rate at the wall (29). Figure 18 shows the influence of finite wall catalytic efficiency
and flight velocity upon the wall heat flux.
In order to compare the different cases, the wall heat flux is presented in reduced
form, divided by its value for a catalytic wall. The faster the wall recombination, i.e.
the more catalytic the wall, the higher the heat flux as more atoms can recombine
at the wall and release their formation enthalpy. As the velocity increases, the gas is
more and more dissociated outside of the boundary layer so that the energy release
due to atom recombination on a catalytic wall is more and more important. The heat
flux reduction due to a non-catalytic wall so increases with velocity.
For a shuttle re-entry at 7 kms- 1 , a non-catalytic wall leads to a heat reduction
of 70% and the reaction-cured glass used for the space shuttle thermal protection
system, with a recombination rate constant of about 1 ms- 1 , to a reduction of about
50% with respect to a catalytic wall.
An interesting point mentioned by Goulard is the evolution of wall heat flux with
te-nperature on a wall with finite catalytic efficiency. The catalytic rate coefficient
increases with temperature; as the wall temperature increases the wall becomes more
and more catalytic so that the wall heat flux may increase with wall temperature.
Sensitivity study of boundary layers computations to the real gas model for dissociating air
Thanks to computer power increase, it was possible to compute dissociating air boundary layer flows along bodies in the late sixties [8, 10]. A sensitivity study of boundary
layer to the real gas model has recently been performed by EIdem [3, 6, 19, 21] for
the STS-2 flight re-entry.
The five main neutral species (N2' O 2 , NO, N, 0) are accounted for, the chemistry
is described with the fifteen dissociation reactions and the two shuffle reactions, thermodynamical properties are computed with Schafer model [42] and transport properties with Straub model [47]. Computations have been performed over axisymmetric
hyperboloids at zero degree incidence which are supposed to represent the space shuttle windward symmetry line during the STS-2 re-entry [46]. The hyperboloid nose
radius and asymptote half-angle are given in table 1. A constant wall temperature
of 1500 K is used in all the computations presented herein.
Influence of wall catalytic efficiency

The stagnation point wall heat flux evolution

Computational Methods in Hypersonic Aerodynamics 319

TOTAL HEAT TRANSFER


CATAfYTlC WfLl.:

0.4

r.oL,J/ V

" ,"

V'

WAi>

~
OJ I-

o
10

-- - ::---- - -

----

//

--1--/

r1

CONOUCTIVE PART OF HEAT TANS1R


TALY C
TOA
WAL

10 D

J-

10

10 l

10 Z

10-'

10

10z

RECOMBINATION RATE PARAMETER C I

Figure 16: Heat transfer parameter as function of the recombination rate parameter

-<1> P

Nu _

7Re - (h", -

H.)

P",/-L",

(due
')
d;

(from [23])

..
u
z

0.6

--

,RECOMBINATION RATE PARAMETER C, 0

0.5

10--

u 004

I-

<l:

c::

lL.

0.3

Vl
Vl

<l:

,...".,...

>-10. 3

~I-*

0.<:

10'

:E 0.1
0

f--

o
o

,/

0.2

0.4

0.6

-==

~
V
..,/'"V
/

/-V

0.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

TJ
Figure 17: Atom mass fraction profile at the stagnation point on a non-catalytic wall as function of
the recombination rate parameter (from [23])

Figure 18: Reduced heat transfer as function of catalytic recombination rate and flight velocity at
250,000 ft altitude (from [27])

320 Computational Methods in Hypersonic Aerodynamics

f"

0.5

Cil.talytic

0.3

0.2 -

0.1

o.

L-_-:--,L---_ _--'-_ _---'_ _ _ --'------.l _ _ _' - - '


200
~OO
600
800
1000
1,WO
1~00

TiUlc(s)

Figure 19: Stagnation point wall heat flux during STS-2 flight re-entry
0,5

catalytic wall

0,

5,

10,

15 ..

20"

30,
X (H)

Figure 20: Wall heat flux evolution along the body at 71.29 km

Computational Methods in Hypersonic Aerodynamics 321


during the re-entry is plotted on figure 19 for a catalytic wall, a non-catalytic wall
and a wall with finite catalytic efficiency according to Scott data [43]. The agreement
with Goulard results is fair as the maximum wall heat flux is reduced by about 70%
for a non-catalytic wall and 40% for a wall with finite catalytic efficiency.
These reductions remain important during all the re-entry, not only at the stagnation point but also all along the body as shown on figure 20. The wall heat flux
reduction due to a non-catalytic wall is roughly constant all along the body. For
71.29 km altitude, where maximum heat fluxes are encountered, a non-catalytic wall
gives radiative equilibrium temperature 250 K lower than a catalytic wall. (It must
be mentioned that the circles on figure 20 are flight measurements. They must not
be compared with the computations as in the computations the wall temperature was
overestimated and the heat fluxes so underestimated. Moreover, entropy swallowing
effects were not accounted for and the geometry was only an approximate one.)
The atomic nitrogen mass fraction profiles are plotted on figure 21. As thermal
diffusivity is accounted for, the slope of the profile at the wall is not null for a noncatalytic wall. Moreover, chemistry leads to nitrogen recombination in the boundary
layer, but the nitrogen mass fraction does not vary significantly when the wall is noncatalytic. Nevertheless, the decrease of atomic nitrogen mass fraction near the wall
with increasing wall catalytic efficiency is clearly brought into evidence.
For a catalytic wall, the fluid is less dissociated in the wall region as shown on
figure 21 so that the temperature is higher as less energy is stored as species formation
enthalpy as shown on figure 23. According to the state equation (9), the density depends upon both the temperature and the dissociation level. Catalytic walls increase
the temperature but decrease the dissociation level so that the influence on density
profile cannot be estimated a priori. The variation of density with surface catalytic
efficiency is not always the same in the boundary layer but is weak as shown on figure 24. Consequently, the velocity profile is hardly modified by the wall catalytic
efficiency, as shown on figure 22.
Other interesting parameters can be obtained from boundary layer computations
such as the skin friction coefficient to know the viscous contribution to the drag
or the displacement thickness to estimate viscous/in viscid coupling effects. These
parameters are related to the velocity and density profiles as
Cf

=,"'--- = #w,(~) w
2Peu~

"2Peue

0,

=f
0

(1 - ~)

dy

PeUe

Figure 25 shows that, as the velocity profile is unchanged, the skin friction is
hardly affected by the wall catalytic efficiency. The skin friction coefficient is very
large for a laminar boundary layer. This is due both to the low Reynolds number and
to the cold wall temperature.
The influence of wall catalytic efficiency upon the displacement thickness is clear
on figure 26 but cannot be a priori estimated. It must be pointed out that the displacement effect is very weak as the wall temperature is low. The rapid increase of
density in the wall region shown on figure 24 reduces the mass flow loss in the boundary layer. Negative displacement thicknesses can be observed at high altitudes or for
lower and more realistic wall temperatures. Viscous/inviscid interaction should be

322 Computational Methods in Hypersonic Aerodynamics

2.5

"'-

G.O

.I
:~

;!
II
"
II

,/

,,

1.0

//"

,,,.,,,.
0.5
Semi-catalytic

",/

...... / / /Non-calalylic I

~' _ _ _ _ _

o.

o.

0.10

(_: _ _ _ _ _ ...J

0.20

0.'0

0.30

Figure 21: Atomic nitrogen stagnation point profiles at 71.29 km altitude


2.0

1.5

1.0

,
,,

0.5

,
CatAlytic wall

,
,,
,,

//

'
/",/"

...................

...... ' ...

--

/,/

Non-<atalytic wall

u:
0 0
0.20
0.'0
0.60
0.80
1.00
Figure 22: Stagnation point velocity profile at 71.29 km altitude

Computational Methods in Hypersonic Aerodynamics 323

2.0

..
'""
~

1.5

il

!I

,
1.0

,
,,
,, ,
,

,
,,
,,
,
,

Non-catalytic wall

0.5

////
Catalytic wAil

... ,"" ..... '

O.

......

~--~~~----~~----~~

1.00

2.00

3.00

'.00

____~~____~
5.00

6.0C

Temperalure (IOOOK)

Figure 23: Stagnation point temperature profile at 71.29 km altitude


2.0

1.5

0.5

'--

ob.L-------~2~.~OO~------~,~.O~O-=======6~.~0~O------~6~.OO

Figure 24: Stagnation point density profile at 71.29 km altitude

324 Computational Methods in Hypersonic Aerodynamics

1
1
1
1
1
\
\
\

8,

\
\
\

1
\
\

\
\
\

\
\

G,

(.,

\,

Ca.lalytic wall

" "
Non--catalytic wall

2,

Distance (M)

'
0, L-----",SL,-----;;'10",-------,',;1':ci-,-----;2;';'0",----'20;;5',-----O;J 0 IS

Figure 25: Influence of the wall catalytic efficiency upon the skin friction coefficient at 71.29 km
3,0
C,(cm)

2.5 Cata.lytic wall

2,0
Non-cata.Jytic waH

1,5 -

1,0 -

0.5 -

Distance (M)

0,

~----~5-"----~,tOn-,-------,1~15~"------;2f.'0~.-----;2~15~,--------~J'O"

Figure 26: Influence of the wall catalytic efficiency upon the displacement thickness at 71.29 km

Computational Methods in Hypersonic Aerodynamics 325

very weak.
Influence of the chemical reaction rates It has already been stated that the reaction
rate constants are known with some uncertainty.
::;- 0.5

,.'e

Chemica.l equilibrium

--,

...........

....../ ...Catalytic wa.1I ' \

,,

"

,,
,/

.........

/ /

Oe.rtel

~'

,,
,,,
\,

'. \

'. \

'.\

\\

.~

.
.'

~\

0.2

"

Non--cataly1.ic wa.1I

Bortner

........, . \
..............

r', . <. . . . .

0.1

Gardiner

'" ~'"

\\

r>

l'i

/
o.

0:-----'2=-=0'=-0--,,,-:O'=-O--:-:60:-::0--8=-=0'=-0--1=-=0'=-00:----:-:12:-::0-=-0--:'1
'00
Time (.)

Figure 27: Influence of the chemical reaction rate on the stagnation point wall heat flux

Three chemical models have been compared, namely the sets of reaction rate
constants published by Bortner [12], Gardiner [24] and Oertel [36]. Gardiner data
are the same as Baulch ones [1] which are the NBS recommended rate constants [51],
except for the two shuffle reactions. A fourth result is obtained assuming infinite
reaction rate constants, i.e. the flow is at chemical equilibrium.
The variation of the stagnation point wall heat flux prediction with the chemical
reaction rate constants is shown on figure 27. For a catalytic wall, the wall heat flux
does not depend upon the reaction rate as atom recombination energy is transferred to
the wall. A slightly larger wall heat flux is obtained for flow at chemical equilibrium.
This seems at variance with Fay and Riddell results and is due to the fact that
they considered very low wall temperatures so that, at chemical equilibrium, the
recombination occurs far from the wall as shown on figure 15 and a part of the
released heat is not transmitted to the wall. In EIdem's computation, as the wall
temperature is higher, the recombination occurs close to the wall and is more complete
for equilibrium flows, so that both the heat release and the wall heat flux are slightly
increased.
For a non-catalytic wall, a large discrepancy is observed as the wall heat flux
depends only upon the wall temperature gradient, i.e. upon the energy release due
to chemical reactions inside the boundary layer. Oertel's chemical model yields too

326 Computational Methods in Hypersonic Aerodynamics

\
\
\
\
\
I

0.05 -

0" . . .
..................

\'

6---

Chemica.l equilibrium

~-~

CataIY~;~ ~-a;("'-:--:--:-::-:-::-:-::-:-::-::-===~O~-:-:_-.!cfl2.._ _ __

.~

..

"~ .~...-: ..~~..~-=. =:-:..;;._=-

Non-catalytic wall

Oertel

,------:='------:-,'0'-,.--_=-_~_.'~~~~..:::7:-: .. ~t=" .. =:- ~=-.--=-:~i-=:~=-:..::=:..:~ortner

o.

5',

10 ,

20 ,

15 "

25 ,

:J 0 .cardiner
Di,tance (M)

Figure 28: Influence of the chemical reaction rate on the wall heat flux at 85.74 km altitude

0.5

~
~

O.l.

~
O,:J

0.2 Chemical equilibrium

\ \,
,

I
"

\,,'
\'

......~

o ;.-__
I,

/Catalytic wall

.. ___-l):;-

--:~~~~-7Non-eatalytic wall

O.

cP

::::O~~;:;-O:o=;e~=-t:::~l-";"':;-~==~===~~----,Bortner

...wGardiner
L---------~5~.---------1~O~.--------~1~5-.------~2~0~.--------~2~5-.------~:J0.
Distance (M)

Figure 29: Influence of the chemical reaction rate on the wall heat flux at 71.29 km altitude

Computational Methods in Hypersonic Aerodynamics 327

fast dissociation rates for oxygen and hence a faster recombination rate in the cold
region close to the wall. A higher wall heat flux is so predicted. Gardiner and Bortner
models, which are within the present uncertainty range, lead to a 12% difference in
wall heat flux, i.e. roughly a 3% difference for the radiative equilibrium temperature
or about 40 K.
The influence of the chemical rate constant on the wall heat flux predictions along
the shuttle centerline is shown on figures 28 and 29 for two points along the re-entry
trajectory. During all the re-entry, the wall heat flux on the shuttle centerline does not
depend upon the chemical model if the wall is catalytic. The wall heat flux predicted
for flow at chemical equilibrium is again slightly larger than the one for a catalytic
wall. The argument presented for the stagnation point still holds. For a non-catalytic
wall, the sensitivity of the wall heat flux prediction to the chemical model reflects the
variations of flow conditions during re-entry:
At higher altitudes (figures 27 and 28), the density is very low, even behind the
shock wave. The chemical time scale is very long and the flow is almost frozen.
The results are then hardly sensitive to the chemical model.
When the altitude decreases, the density increases very rapidly as shown on
table 1. The chemical time scale decreases rapidly so that the flow is in chemical
non-equilibrium. Oertel's model gives larger oxygen recombination and higher
wall heat flux. Some discrepancies are observed between Bortner and Gardiner
models.
At lower altitudes (figures 27 and 29) the density still increases and the flow
is still in chemical non-equilibrium. The discrepancies are amplified. Oertel's
model gets close to catalytic wall results as it tends to recombine all the oxygen.
Bortner's model predicts a wall heat flux roughly 40% larger than Gardiner's
model, i.e. about 70 K discrepancy for the radiative equilibrium wall temperature
on the rear part of the shuttle.
At lower altitudes, the velocity has decreased while the density still increases so
that the flow gets closer to chemical equilibrium. As the velocity is lower, only
oxygen dissociates now. Oertel's model gives predictions similar to equilibrium
flow. Bortner and Gardiner models agree and still show a wall catalytic efficiency
effect, i.e. chemical equilibrium is not yet reached.
The study of the influence of chemical reaction rate on the other boundary layer
parameters shows that the velocity profile is not modified while the mass fraction
and temperature profile depend upon the reaction rate, so that the evolution of the
density profile cannot be a priori predicted. Consequently the skin friction coefficient
is not significantly modified by the chemical model while large variations of the displacement thickness, about a factor of two, are observed but the displacement effect
remains small.
Influence of the wall temperature As foreseeable, it is observed that the wall heat
flux is higher when the temperature is lower. The effect of wall temperature is more
important for a non-catalytic wall than for a fully catalytic wall. The effect is more
important when the altitude decreases.

328 Computational Methods in Hypersonic Aerodynamics


These results are consistent with the idea that the wall heat flux is proportional
to the difference between the wall enthalpy hw and the recovery enthalpy hi:

For hypersonic flows, the friction enthalpy hi can be approximated by the inviscid
flow stagnation enthalpy He. The effect of the wall temperature is more important
when the difference hi - hw is smaller. The difference hi - hw decreases when the
velocity decreases as the friction enthalpy becomes smaller. The difference hi - hw
is also smaller when the wall enthalpy is larger: for a given wall temperature, the
wall enthalpy is larger when the dissociation is more advanced, which is the case for
a non-catalytic wall.
Model reduction
In order to decrease the computational cost, it is important to identify the key features
of the real gas model and to discard unimportant phenomena which may require large
computational time.
A systematic study of model simplifications from the analysis of "complete" model
computations has been performed by EIdem for boundary layer flows on the STS-2
re-entry [3,6, 19,21].
Concerning chemical models, a good approximation to represent all the process
with a reduced set of chemical reactions is the Zeld'ovich model

o +0 +

r= NO+

O2

N2

At low altitudes, the prediction can be improved by taking into account two more
oxygen dissociation reactions

as only oxygen dissociates at the end of the re-entry due to the decrease of the velocity. These results are in agreement with previous results obtained by Blottner [9]
who brought into evidence the major role of the oxygen dissociation and the shuffle
reactions. To get a perfect agreement with the complete computations all over the
re-entry trajectory, five more reactions of nitrogen and nitrogen monoxide dissociation
are needed.

N2 +N2
N2 +N
NO+N2
NO+N
NO+O

-->.
.,.-->.

.,.-->.
.,.-->.

,..-

-->.
,..-

N+N+N2
N+N+N
N+0+N2
N+N+O
N+O+O

The seven other reactions play no role and can be neglected.


Concerning the transport model, the thermal diffusion is a second order effect and
can be neglected. Moreover, the Prandtl number remains quite constant in all the
flow during the whole re-entry so that a constant Prandtl number about 0.72 can be

Computational Methods in Hypersonic Aerodynamics 329


assumed without modifying model predictions. At last, analysis of the species diffusion
shows that it can be roughly modelled by a Fick law but that the Lewis number is not
the standard 1.4 value but closer to 1, in agreement with Cohen results [18] shown on
figure 14. A value of 1.2 gives fair predictions all over the STS-2 re-entry trajectory.
Concerning the thermodynamical model, polynomials fits to thermodynamic functions [26] give the same results as the use of more complex (and computationally time
consuming since they involve exponentials) expressions for the energy partition function to compute thermodynamical functions.
Influence of ionization in re-entry flows
This topic is addressed in [5] in which boundary layer computations with and without
ionized species and ionization chemistry are compared. For shuttle re-entry flows, i.e.
for velocity about 7000 m S-l, ionization is weak. Only about one percent of nitrogen
monoxide is ionized so that ionization hardly affects the energy budget. Ionization
can therefore be neglected for such flows.
THREE-DIMENSIONAL BOUNDARY LAYERS
Self-similar solutions
Stagnation point
Since Howarth's works [32, 35], it is customary to study threedimensional stagnation points in a cartesian coordinate system fixed in a plane osculating the stagnation point. In the vicinity of the stagnation point, with the axis
origin at the stagnation point, the inviscid pressure distribution can be expressed by
the following expansion:

i.e. as the pressure is maximum at the stagnation point, the iso-pressure contours
near the stagnation point are ellipses of identical axes. Taking these ellipses axes as
coordinate axes, the pressure distribution reads:

with a = b for an axisymmetric stagnation point.


On the other hand, the momentum equations for the inviscid flow at the wall (or
for the outer edge of the boundary layer) reduces, in the cartesian coordinate system,
to:

ap

ax

ap
az

(75)

(76)

so that, by taking derivatives of these two equations respectively with respect to X


and Z, the velocity gradients at the stagnation point are obtained as:

aw

az

fib
vr;

(77)

330 Computational Methods in Hypersonic Aerodynamics


There exist two orthogonal directions along which the inviscid flow is radial; these
directions and the associated velocity gradients can be determined from the knowledge
of the pressure field.
An extension of the Levy-Lees-Dorodnitsyn space variable transformation (6S,
69) which takes advantage of the existence of two principal directions and also reduces
to the standard form for two-dimensional flows has been looked for as:
'TJ

= o:(X, Z)

loy pdy

(7S)

The self-similarity conditions together with the identity with the Levy-Lees-Dorodnitsyn transformation for two-dimensional, plane and axisymmetric, flows impose [5, 21]:

au

0:=

aw

fiX+az
pw/-lw

(79)

The boundary layer equations can then be reduced to a set of ordinary differential
equations at the stagnation point, using this space transformation and looking for selfsimilar solutions for the variables:
I

9 =We

H
s=He

(SO)

For axisymmetric stagnation points, the wall heat flux is proportional to the
square root of the stagnation point velocity gradient, or if the pressure distribution is
given by a Newtonian law, to the inverse of the square root of the nose radius. The
proportionality constant accounts for gas properties, external flow dissociation level
and wall catalytic efficiency (70 to 73).
A parametric study of three-dimensional stagnation points has been performed in
order to try to extend these formulae to three-dimensional flows. The set of results
can be fitted with the simple formulae [21]:
(Sl)
or for a Newtonian pressure distribution:

(S2)
where Rl and R2 are the principal curvature radii at the stagnation point. Of course,
the proportionality constants are the same as for axisymmetric flows.
Attachment line Important heat loads can be encountered along the attachmentlines
on the leading edges of wings or winglets where the curvature radii are small.

Computational Methods in Hypersonic Aerodynamics 331

'<7~--"

U
oo

This problem can be tackled by modelling the leading edge as an infinite swept
cylinder. Self-similar solutions can be used to study the flow along the attachment
line. The inviscid flow velocity along the attachment line We is constant while there
exists a velocity gradient normal to the attachment line (u e = kx). A space change
of variable analog to the Levy-Lees-Dorodnitsyn one is used again
Ue fY
17 = ~ io pdy

(=z

(83)

and self-similarity along the attachment line is sought for the dimensionless boundary
layer profiles

f' = ~

= !!!...

s=

ZI

CI
(84)
C1e
Self-similar solutions for an infinite swept wing attachment line have already been
studied for perfect gas flows [34,41]. For incompressible flows as well as for hypersonic
flows with the Newtonian approximation, the velocity gradient k is proportional to the
cosine of the sweep angle. For incompressible flows, the solution of the self-similarity
equations is not affected by the sweep angle and thus the wall heat flux is proportional
to the square root of the sweep angle cosine [34].
Ue

g'

We

He

(85)

For supersonic and hypersonic flows, the viscous dissipation term in the energy equation has a growing importance as the sweep angle increases. Michel's results [34] for
high Mach numbers (M > 7) have been fitted by [25]:
qt
=
( qt)",=o

COS 3 / 2

'ljJ

(86)

An example of the wall heat flux evolution with the sweep angle for flows with real
gas effects is given on figure 30 for a cylinder of 0.3 meter radius. For a catalytic wall,
the cos3 / 2 'ljJ evolution is still observed while for a non-catalytic wall, an unexpected
behaviour is observed.
The same heat flux predictions for perfect gas flows and chemical non-equilibrium
flows on catalytic walls has already been evidenced in two-dimensional flows [3, 5, 19,
21]. It corresponds to the fact that no energy can be stored as formation energy of
the atoms at the wall so that all the available energy is transmitted to the wall.

332 Computational Methods in Hypersonic Aerodynamics

600

.... --.

500

~ ......
"-,

' " Catalytic wall

400
300

200

.---

.-

'"

Non catalytic wall

100

10

20

30

40

50

'"~ i'...

'-.

60
70
80
90
Sweep angle (degrees)

Figure 30: Wall heat flux along an infinite swept cylinder attachment line

For zero degree sweep, the ratio between the heat fluxes on catalytic and noncatalytic walls is similar to the one previously obtained for axisymmetric or threedimensional stagnation points at the same altitude. For very large sweep, close to
ninety degrees, as the shock is parallel to the cylinder, the shock is weak and little
temperature increase and real gas effects occur behind the shock so that no wall
catalytic efficiency influence is observed.
Let us try to understand the increase of the wall heat flux on a non-catalytic wall
for intermediate sweep angles. We assume that the heat flux is proportional to the
stagnation enthalpy variation through the boundary layer divided by the boundary
layer thickness. As the boundary layer thickness is unaffected by the wall catalytic
efficiency,
He - hwnon-catalytic (
(87)
(qt)non-catalytic = H _ h
.
qt)catalytic
e
wcatalytJc
The different behaviour of the wall heat flux on catalytic and non-catalytic walls must
reflect different evolutions of the wall enthalpy.
On a catalytic wall, the fluid recombines at the wall and behaves locally as a
perfect gas: the wall enthalpy is proportional to the wall temperature and does not
depend upon the sweep angle.
On a non-catalytic wall, the fluid can remain dissociated at the wall. The wall
enthalpy then accounts for the enthalpy formation of atoms and nitogen monoxide
and is larger than for a perfect gas. For altitudes higher than 60 km, the flow is
roughly chemically frozen so that the species mass fractions are constant throughout
the boundary layer provided the wall is non-catalytic. The dissociation level is thus

Computational Methods in Hypersonic Aerodynamics 333


the same outside of the boundary layer and at the wall; the wall enthalpy is directly
linked to the dissociation level of the external flow. When the sweep angle increases,
the velocity along the attachment line increases and the flow enthalpy outside of the
boundary layer decreases. The pressure also decreases when the sweep angle increases.
If the external flow is assumed to be at chemical equilibrium, the enthalpy variation is
the leading term, the dissociation level of the external flow decreases. Consequently,
the wall enthalpy decreases as the sweep angle increases. The behaviour of the wall
heat flux on a non-catalytic wall is thus the combination of the COS 3 / 2 'IjJ decrease of
the catalytic wall heat flux and of the wall enthalpy decrease.
Sensitivity study
The sensitivity study of boundary layer computations to the real gas model has
been extended to three-dimensional flows [5, 11]. The phenomena observed in twodimensional flows, namely the importance of wall catalytic efficiency, of the chemical
model or of the wall temperature, together with the evolution of the wall heat flux
along the re-entry trajectory are retrieved in three-dimensional boundary layers.
Examples of three-dimensional boundary layer computations over a double ellipsoid or an approximate geometry of the forward part of the space shuttle can be found
in [4, 5, 11]. An example of wall heat flux distribution on an approximate geometry of
the forward part of the space shuttle is displayed on figures 31 and 32 for a catalytic
and a non-catalytic wall. The iso-contours look similar but the levels are changed
from one figure to the other, which evidences a similar heat flux reduction due to the
non-catalytic wall as for two-dimensional flows.
An interesting point is that the maximum heat flux is no longer at the sta.gnation
point but by its sides. This is due to the fact that, near the stagnation point, the wall
heat flux is directly related to the velocity gradient. Since, at the stagnation point,
the wall heat flux depends upon the mean of the velocity gradient in the two principal
directions (81), it may increase in the direction of the maximum velocity gradient.
CONCLUSIONS
Real gas effects deeply affect the evolution of the boundary layer on a space plane
during re-entry. They are due to the transformation of kinetic energy into heat downstream of the shock wave which leads to chemical reactions and dissociation (even
ionization) of the flow. Thermal equilibrium can be assumed in the boundary layer
so that the analysis is simplified.
The governing equations for a gas mixture have been presented. Species equations
bring into evidence various behaviours according to the relative order of magnitude
of the advection, diffusion and chemistry terms. All the transport coefficients which
appear in the governing equations must be determined from gas kinetic theory. The
chemical model has to be taken from experimental data. While species thermodynamic properties are known with good accuracy, this is not the case for chemical
reaction rates or for the mixture transport properties. For boundary layer flows, wall
catalytic efficiency appears in the boundary conditions.
A dissociated diatomic gas model has been introduced to illustrate real gas effects.
Fay and Riddell stagnation point solutions bring into evidence the influence of wall

334 Computational Methods in Hypersonic Aerodynamics

kW/m2
_

> 450
400450
350400
300350

_
_
_

250 300
200250
150 200
100 150
50 100
<
50

Figure 31: Wall heat flux distribution at 71.29 km - catalytic wall

kW/m2
_

_
_
_
_

l1li
_

J;ltiiifrn
1;:';",1
I'~H

Figure 32: Wall heat flux distribution at 71.29 km - non-catalytic wall

>

135

120135
105120
90105
75 90
60 75
45 60
30 45
15 30
<
15

Computational Methods in Hypersonic Aerodynamics 335


catalyticity and chemistry kinetics on the wall heat flux.
More realistic models for dissociated air can be used nowadays. Calculations are
performed for the STS-2 re-entry trajectory. They bring into evidence the major
role of wall catalytic efficiency and the influence of the incertainties about chemical
reaction rates. The influence of wall temperature is also analysed. Moreover, model
simplifications are proposed in order to save computation time: the set of chemical
reactions can be reduced to a restricted set of predominant reactions, transport modelling can be drastically simplified and thermodynamic functions can be fitted with
polynomials.
Three-dimensional boundary layer computations are needed for realistic shapes
and practical problems. For three-dimensional flows, self-similar solutions can be
used to compute the flow at the stagnation point or along the leading edges, at least
when they are close to infinite swept cylinders. A surprising increase of the wall
heat flux with sweep angle for a non-catalytic wall has been evidenced and explained.
The sensitivity of the flow predictions to the real gas models is the same in threedimensional as in two-dimensional flows.
A major problem is the validation of such computational approaches. Flight data
on shuttle re-entry [46] only give wall temperature and heat fluxes on a complicated
three-dimensional body. So experiments are needed to be able to validate mainly the
chemical reaction rates but also the transport coefficients or the numerics in realistic
situations.
The authors wish to acknowledge C. EIdem, S. Bonnet, J. P. Brazier and S. Grunwald for their important contribution to the study of real gas effects in boundary layer
flows.

336 Computational Methods in Hypersonic Aerodynamics

References
[1] J. D. Anderson, Jr. Hypersonic and High Temperature Gas Dynamics. Mc GrawHill Book Company, 1989.
[2] S. Arrhenius. Uber die Reaktiongeschwindigkeit bei der Inversion von Rohrzucker
durch Saiiren. Z. Physic. Chem., 4(226), 1889.
[3] B. Aupoix. An introduction to real gas effects. In Special Course on Aerothermodynamics of Hypersonic Vehicles. AGARD-FDP VKI Lecture Series - AGARD
Report 761, 10 May - 3 June 1988.
[4] B. Aupoix, S. Bonnet, C. Gleyzes, and J. Cousteix. Calculation of threedimensional boundary layers including hypersonic flows. In Fourth Symposium
on Numerical and Physical Aspects of Aerodynamic Flows - Long Beach, 14-19
January 1989.
[5] B. Aupoix and J. Cousteix. Real gas effects in hypersonic laminar boundary
layers for reentry flows. In Second Joint Europe-US Short Course on Hypersonics
- Colorado Springs, 16-20 January 1989.
[6] B. Aupoix, C. EIdem, and J. Cousteix. Couche limite laminaire hypersonique:
Etude parametrique de la representation des effets de gaz reel. In Aerodynamics
of Hypersonic Lifting Vehicles - AGARD-CP-428, 6-9 April 1987.
[7] D. 1. Baulch, D. D. Drysdale, and D. G. Horne. Evaluated Kinetic Data for
High Temperature Reactions - Volume 2 Homogeneous Gas Phase Reactions of
the H2 - N2 - O2 System. London Butterworths, 1973.
[8] F. G. Blottner. Chemical nonequilibrium boundary layer.
2(2):232-240, February 1964.

AIAA Journal,

[9] F. G. Blottner. Nonequilibrium laminar boundary-layer flow of ionized air. AIAA


Journal, 2(11):1291-1297, November 1964.
[10] F. G. Blottner. Electron number density distribution in the laminar air boundary
layer on sharp cones. AIAA Journal, 7(6):1064-1069, June 1969.
[11] S. Bonnet. Couches limites laminaires tridimensionnelles avec effets de dissociation. PhD thesis, Ecole Nationale Superieure de l'Aeronautique et de l'Espace,
Toulouse, 17 Juin 1988.
[12] M. H. Bortner. A review of rate constants of selected reactions of interest in
re-entry flow fields in the atmosphere. Technical Report 484, National Bureau of
Standard, May 1969.
[13] D. Boyer. Species composition measurements in non equilibrium high speed flows.
In Special Course on Aerothermodynamics of Hypersonic Vehicles. AGARD-FDP
VKI Lecture Series - AGARD Report 761, 10 May - 3 June 1988.

Computational Methods in Hypersonic Aerodynamics 337

[14J J. P. Brazier, B. Aupoix, and J. Cousteix. Second-order effects in hypersonic laminar boundary layers. In Computational Methods in Hypersonic Aerodynamics.
Computational Mechanical Publications, Ashurst Lodge, Ashurst, Southampton
S04 2AA, United Kingdom, 1991.
[15J R. Brun. Transport et relaxation dans les ecoulements gazeux. Masson - Physique
fondamentale et appliquee, 1986.
[16J M. L. Carnicom. Reaction rate for high-temperature air with carbon and sodium
impurities. Technical Report SC-R-68-1799, Sandia Laboratories, May 1968.
[17J S. Chapman and T. G. Cowling. Mathematical Theory of Non Uniform Gases.
Cambridge University Press, 1939.
[18J N. B. Cohen. Correlation formulas and tables of density and some transport
properties of equilibrium dissociating air for use in solutions of the boundarylayer equations. Technical Report TN D-194, NASA, February 1960.
[19J J. Cousteix and B. Aupoix. Calculation of hypersonic laminar boundary layers. In
First Joint Europe-US Short Course on Hypersonics - Paris. Birkauser, Boston,
7-11 December 1987.
[20J W. H. Dorrance. Viscous Hypersonic Flow. Mc Graw-Hill Book Company, 1962.
[21J C. EIdem. Couches limites hypersoniques avec effets de dissociation. PhD thesis, Ecole Nationale Superieure de l'Aeronautique et de l'Espace, Toulouse, 14
Decembre 1987.
[22J P. Fauchais. Measurements techniques in high temperature gases: Temperatures,
equilibrium conditions, species densities, flow velocity. In First Joint Europe-US
Short Course on Hypersonics - Paris. Birkauser, Boston, 7-11 December 1987.
[23J J. A. Fay and F. R. Riddell. Theory of stagnation point heating in dissociated
air. Journal of the Aeronautical Sciences, 25(2):73-85, 121, February 1966.
[24J W. C. Gardiner, Jr. Combustion Chemistry. Springer-Verlag, 1984.
[25J J. P. Gilly, L. Rosenthal, and Y. Semezis.
Gauthier-Villars, 1970.

Aerodynamique Hypersonique.

[26J S. Gordon and B. J. Mc Bride. Computer program for the calculation of complex
equilibrium compositions, rocket performance, incident and reflected shocks and
Chapman-Jouquet detonations. Technical Report SP 273, NASA, 1971.
[27J R. Goulard. On catalytic recombination rates in hypersonic stagnation point heat
transfer. Jet Propulsion, pages 737-745, November 1958.
[28J R. N. Gupta, J. M. Yos, R. A. Thompson, and K. P. Lee. A review of reaction
rates and thermodynamic and transport properties for an II-species air model for
chemical and thermal nonequilibrium calculations to 30000 K. Technical Report
Reference Publication 1232, NASA, August 1990.
[29J J. Heicklen. Gas-phase chemistry of re-entry. AIAA Journal, 5(1):4-15,1967.

338 Computational Methods in Hypersonic Aerodynamics


[30] E. H. Hirschel and W. Kordulla. Shear Flow in Surface-Oriented Coordinates, volume 4 of Notes on Numerical Fluid Mechanics. Vieweg, BraunschweigJWiesbaden, 1981.
[31] J. O. Hirschfelder, C. F. Curtiss, and R. B. Bird. Molecular Theory of Gases and
Liquids. John Wiley and Sons, 1954.
[32] 1. Howarth. The boundary layer in three-dimensional flow. II: The flow near a
stagnation point. Phil. Mag., 7(1433-1440), 1951.
[33] S. W. Kang and M. G. Dunn. Theoretical and measured electron density distributions for the RAM vehicle at high altitudes. AIAA Paper 72-689, 1972.
[34] R. Michel and N. Due-Lam. Frottement et transfert de chaleur turbulents en
ecoulements bi et tridimensionnels. In 11th International Congress of Applied
Mechanics, Munich, 1964. Springer-Verlag.
[35] F. K. Moore. Theory of Laminar Flows. High Speed Aerodynamics and Jet
Propulsion. Princeton University Press, 1964.
[36] H. Oertel. Sto{3rohre. Springer-Verlag, 1966.
[37] C. Park. Problems of rate chemistry in the flight regimes of aeroassisted orbital transfer vehicles. AIAA Paper 84-1730 19 th Thermophysics Conference Snowmass, Colorado, July 25-28 1984.
[38] C. Park. On convergence of computation of chemically reacting flows. AIAA
Paper 85-247 23 rd Aerospace Sciences Meeting - Reno, Nevada, January 14-17
1985.
[39] C. Park. Nonequilibrium Hypersonic Aerothermodynamics. John Wiley & Sons,
1990.
[40] C. Park and J. P. Meenes. Odd nitrogen production by meteroids. Journal of
Geophysical Research Series C, 83(8):4029-4035, 1978.
[41] E. Reshotko and I. E. Beckwith. Compressible laminar boundary layer over a
yawed infinite cylinder with heat transfer and arbitrary Prandtl number. Report
1379, NACA, 1958.
[42] K. Schafer. Statistische Theorie der Materie. Vandenhoek Gottingen, 1960.
[43] C. D. Scott. Catalytic recombination of nitrogen and oxygen on high-temperature
reusable surface insulation. AIAA Paper 80-1477 15 th Thermophysics Conference
- Snowmass, Colorado, July 14-16 1980.

[44] C. D. Scott. The effects of thermochemistry, nonequilibrium and surface catalysis


in the design of hypersonic vehicles. In First Joint Europe-US Short Course on
Hypersonics - Paris. Birkauser, Boston, 7-11 December 1987.
[45] C. D. Scott. Wall catalytic recombination and boundary conditions in nonequilibrium hypersonic flows - with applications. In Third Joint Europe-US Short
Course on Hypersonics - Aachen. GAMNI/SMAI, 1-5 October 1990.

Computational Methods in Hypersonic Aerodynamics 339


[46J J. Shinn, J. N. Moss, and A. L. Simmonds. Viscous shock-layer heating analysis
for the shuttle windward symmetry plane with surface finite catalytic recombination rates. AlA A Paper 82-0842 3rd AlAA/ AS ME Joint Thermophysics, Fluids,
Plasma and Heat Transfer Conference - Saint Louis, Missouri, June 7-1 1982.
[47J D. Straub. Exakte Gleichungen fur die Transportkoeffizienten eines Funfkomponentengemisches als Modellgas Dissozierter Luft. Technical Report FB 72-34,
DLR,1972.
[48J M. Van Dyke. Higher approximations in boundary-layer theory - part 1: General
analysis. Journal of Fluid Mechanics, 14:161-177, 1962.
[49J W. G. Vincenti and C. H. Kruger, Jr. Introduction to Physical Gas Dynamics.
Robert E. Krieger Publishing Company, Malabar, Florida, 1965.
[50J F. Wecken. Donnees concernant la cinetique des reactions dans l'air entre
500K et 10000K. Note de Documentation D 4/68, lnstitut Franco-Allemand
de Recherches de Saint-Louis, 3 Septembre 1968.
[51J F. Westley. Table of recommended rate constants for chemical reactions occuring
in combustion. Technical Report NSRDS-NBS 67, National Bureau of Standard,
April 1980.

Chapter 10:
Flow Analysis and Design Optimization Methods
for Nozzle-Afterbody of a Hypersonic Vehicle
O. Baysal
Old Dominion University, Mechanical Engineering and
Mechanics Department, Norfolk, Virginia 23529, USA
1. INTRODUCTION
The recent resurgence of interest in hypersonic aerodynamics has come about largely
in part due to the development of hypersonic vehicles, such as, the National Aerospace
Plane (NASP). The design of this type of aircraft will rely heavily on the use of
computational fluid dynamics, since the operating conditions prohibit the use of most of
the conventional experimental facilities to obtain the required data for design analysis.
One of the major design tasks involved in the development of a hypersonic airbreathing aircraft is the integration of the engine and the airframe. This is necessary
in order to reduce excessive drag and weight due to the Mach numbers at which the
aircraft will be traveling. The high pressure combustion products are expanded through
the combustor exit nozzle and over the airframe afterbody configuration (Fig. 1). The
overall propulsive efficiency of the nozzle is determined, to a large extent, by the
exhaust plume flow over this afterbody section.
The design and testing of a scramjet nozzle-afterbody section using actual engine
combustion products is impractical in a conventional wind tunnel. The actual chemistry
and high total enthalpy levels of the exhaust products would be quite difficult to
match in a scaled test section. However, several alternatives do exist. A simulant
gas can be substituted for the actual combustion products, provided that dynamic and
thermodynamic similitude are enforced. Perhaps a more economical alternative would
be to do the preliminary design analysis using computational fluid dynamics (CFD).
Since there is currently very little experimental data for very high Mach number flows,
some means of calibrating and validating these CFD codes must be achieved before
they can be used with complete confidence in this design process.
In the 1970's, a study was undertaken to develop an experimental cold gas simulation technique for scramjet exhaust flows [1]. It was determined that in addition to the
usual nondimensional similitude parameter requirements for inviscid flows (i.e. Mach

342 Computational Methods in Hypersonic Aerodynamics


numbers, pressure ratios, temperature ratios, etc.), that the ratio of specific heats (,)
of the combustion products must also be matched by the simulant gases. It was also
determined in this study that the surface pressures were relatively insensitive to small
changes in the thermodynamic properties of the gases, but were very sensitive to flow
perturbations caused by the nozzle geometry.
An extension of this work was carried out recently [2, 3]. A wind tunnel model
of a single-module scramjet nozzle-afterbody configuration was constructed for testing
(Fig. 2). The simulant gas mixture was fed into a high pressure plenum chamber via a
mounting strut. The gas in this plenum chamber was expanded through a convergingdiverging supersonic nozzle to approximately Mach 1.7 at the combustor exit plane,
where it was further expanded over the nozzle-afterbody section of the model. This
supersonic exhaust flow also encountered a hypersonic (Mach 6) freestream air flow,
through which mixing occurred in a free shear layer containing additional expansions
and shock waves. A removable tapered flow fence was used to simulate a quasi twodimensional flow. When this fence was removed, the nozzle flow also mixed with the
hypersonic free stream in the lateral direction through a span wise expansion, causing the
flow to become fully three-dimensional. Experimental data was obtained for a scaled
scramjet nozzle-afterbody flowfield using both air and a Freon!Argon mixture as the
simulant gas. Static pressures were measured on the afterbody surface, for both twodimensional and three-dimensional flows, with various nozzle-afterbody geometries.
Also, by using a flow rake specifically designed for this purpose, the off-surface flow
was surveyed to obtain the pitot pressures. The data obtained from these experiments
were used to compare with the present computational results.
The design and analysis processes for this type of nozzle-afterbody section is
complex due to the fact that many additional parameters must be considered, in addition
to those which must be accounted for in conventional nozzles. This particular nozzle is
highly asymmetric, and consists of an internal and an external portion. The forces and
moments generated by most conventional nozzles can be determined by analyzing the
flow up to the nozzle exit plane only. In this particular case, the analysis must extend
further downstream due to the fact that the lower aft portion of the aircraft forms the
external portion of the nozzle. The flow over this afterbody region is expected to have
a dramatic effect on the thrust vector and pitching moment generated by the engine
module.
In the present study, a simplified configuration (Fig. 3) is assumed to model the
single-module scramjet nozzle-afterbody. A rectangular duct precedes the internal
nozzle. The external part of the nozzle is bounded by a ramp, a side ramp and a
vertical reflection plate. The external hypersonic flow is initially over a double-comer
formed by the reflection plate, the top surface of the nozzle, the exterior of the nozzle
sidewall, and a side flat plate. Both of the flows expand over the 20 ramp and the side

Computational Methods in Hypersonic Aerodynamics 343


ramp. The supersonic jet expands in the axial, the nonnal, and the span wise directions
after it clears the nozzle exit plane. A three-dimensional shear layer structure fonns
between these coflowing turbulent streams which are at different speeds.
In this chapter, the computational methods developed for the flow analysis and the
design of the aforementioned nozzle-afterbody are discussed. The three-dimensional
analysis method for the air-air (simulant gas is air) flow is given in the next section.
A two-dimensional, multispecies flow model is developed for the flow of Argon-Freon
mixing with air, which is explained in Section 3. The results of the flow analyses are
presented in Section 4. Further details of these flow analysis methods and the results
obtained using them may be found in [4-9]. The last two sections are dedicated to the
design optimization of the nozzle-afterbody. The methodology is described in Section
5 and some sample results are included in Section 6. More comprehensive discussion
of this design optimization method may be found in [10-13].
2. ANALYSIS METHOD FOR AIR-AIR FLOW
The conservative fonn of the nondimensional, unsteady, compressible, Reynoldsaveraged, complete Navier-Stokes equations are written below in generalized curvilinear
coordinates,

aQ + aem
a (-E
at
where

Q=

-) m

- Ev

= OJ

m = 1,2,3
T

(2.1)

(2.2)

[p,pUl,PU2,pu3,pe] jJ

The symbols t, p, Ui, e denote the time, the density, the Cartesian velocity components
and the total energy, respectively. The inviscid fluxes, viscous fluxes, and the coordinate
transfonnation jacobian are denoted by E, Ev, and J, respectively. The state equations
are written assuming air to be a perfect gas. Molecular viscosity is calculated using the
Sutherland's law and the Stoke's hypothesis.
A finite volume differencing is fonnulated by integrating the conservation equations
over a stationary control volume,

%t

JJJ + JJE
QdO

ndS

=0

(2.3)

where n is the unit nonnal vector pointing outward from the surface S bounding
the volume O. This implicit and second-order accurate method is described in [14,
15]. The flux-difference splitting [16] is used to construct the upwind differences for
the convective and pressure tenns. Spatial derivatives are written conservatively as
flux balances across the cell. The Roe-averaged cell interface values of fluxes are
evaluated after a state variable interpolation where the primitive variables are used.

344 Computational Methods in Hypersonic Aerodynamics


The diffusion terms are centrally differenced. Spatial approximate factorization and
Euler backward integration after linearization in time, result in the solution through
5 x 5 block-tridiagonal matrix inversions in each of the three directions.
The modeling of the stresses resulting from the Reynolds averaging of the governing
equations is complicated by the fact that several length scales exist which control the
generation, transport, and dissipation of turbulent kinetic energy. Therefore, the standard
two-layer algebraic turbulence model of Baldwin and Lomax [17] is modified and used
herein. It is based on the Boussinesq approximation of modeling the Reynolds stresses
by an eddy viscosity, c. That is, the Reynolds stresses and heat fluxes are assumed
proportional to the laminar stress tensor with the coefficient of proportionality defined
as the eddy viscosity coefficient.
Three specific modifications have been made to the standard Baldwin-Lomax model
to account for: (a) vortex-boundary layer interaction and separation, (b) presence of
multiple walls, and (c) turbulent memory effects in addition to the local equilibrium for
the shear layer. The details of these modifications are given in [9].
The computational domain (11.1 in by 8.1 in by 6.6 in) consists of the region
above the cowl and to the right of the side wall where the flow is hypersonic, and another
region bounded by the lower surface of the cowl and the ramp, where the supersonic
internal nozzle flow expands (Fig. 3). The global grid, which consists of 808,848 cells,
is block-structured with eight subdomains in order to ease the grid generation [8, 9]. The
grid lines are contiguous across the block interfaces, where the solutions are matched
with flux conservation. The step sizes normal to the wall vary in the range of 10-5 to
10-4 with respect to the throat height. The grid is also longitudinally clustered around
the corners inside the nozzle, where the expansions occur. The step sizes for the shear
layer vary from 10-4 to 10-3 with respect to the ramp length in the (-direction.
The upstream boundaries for the external and internal regions require specifying
a viscous, double-corner flow (Fig. 4) profile and a viscous, duct (Fig. 5) profile,
respectively [8, 9]. Generating such profiles requires solving the three-dimensional
compressible Navier-Stokes equations. The boundary layer thickness of the final crossplane profile of the duct flow, which is used as the upstream boundary condition for the
nozzle, is approximately 0.072 in on all four walls (Fig. 4). In addition to the boundary
layer growth on the walls and in the corner regions of the external double-corner, the
interaction of the two co-flowing hypersonic flows are computationally captured (Fig. 5).
No slip, impermeability, adiabatic, and zero-normal-gradient of pressure conditions
are imposed on all solid surfaces. First-order extrapolation for the conserved variables
are used at the downstream boundary. The outer boundary conditions are specified after

Computational Methods in Hypersonic Aerodynamics 345


checking the sign of the nonnal contravariant velocity; extrapolation is used if the flow
is outward and freestream values are used if the flow is inward.
The solution is obtained on two coarser level grids, and finally the finest grid, in
an attempt to overcome the initial numerical transients. This approach is commonly
known as mesh sequencing [14]. The residual and the nonnal force histories are used to
detennine the solution convergence. The convergence is deemed to be achieved when
the residual is decreased by four orders of magnitude. An examination of the nonnal
force coefficient, CN, reveals an asymptotic approach to a constant value after 1500
work units. A work unit corresponds to the amount of iterations on any combination
of coarse or fine grids, which requires the same amount of computer time necessary to
perfonn one iteration on the finest grid [14]. The solution is tenninated at approximately
2300 work units, in which 300 work units are perfonned on coarser levels. This amounts
to roughly 30 hours on the CRAY-2 of NASA Langley Research Center.
3. ANALYSIS METHOD FOR MULTISPECIES FLOW
This method requires solving more equations than the method for the air flow due
to the multi species gases. Therefore, it is shown here in two-dimensions for brevity and
computational time savings. Extending it to three-dimensions is rather straightforward.
The conservation fonn of the two-dimensional, Reynolds-averaged .Navier-Stokes
equations for unsteady, compressible flows of multispecies fluids is being solved. The
nondimensional indicial fonn (i and j are dummy indices) of these equations in the
Cartesian coordinates is given by
(3.1)

i = 1,2

where

Q = [p,

PUi, pe, pIs] T ;

= 1,2, . .. ,]V -

(3.2)
j = 1,2

(3.3)
The mass fraction and pressure are denoted by I and p, respectively. N is the number
of species and indices r and s indicate species. The expressions for the shear stresses
and the heat flux are given as

Tij =

[(aUi
aXj + aUj)
aXi + A(aUi)]
aXj bij

Re f1

) uXi

.
( f1
aT
qi=-C
p -p +-p ~+DMTEi
r

rt

(3.4)

(3.5)

346 Computational Methods in Hypersonic Aerodynamics


Prandtl, Mach, and Reynolds numbers are denoted by Pr, M, and Re, respectively.
First and second viscosity coefficients are shown by f1 and A. T denotes the temperature
and subscript (t) denotes a turbulent quantity. Cp is the specific heat. In the above
system, all the gases are assumed to be thermally perfect but calorically real gases.
Hence, the enthalpy (h) of each species (8), the total energy, and the pressure can be
expressed as:

(3.6)

e = hsfs -

1
p+ 2"(U
iUi )

(3.7)

PRT(~:)

(3.8)

The enthalpy of formation, universal gas constant, and molecular weight are denoted
by hO, R, and w, respectively. The terms DMT Pi and DMT Ei in Eqs. (3.3) and (3.5)
account for the diffusive mass transfer. The expressions for these terms depend on the
utilized diffusion model. In case of using Fick's law, these terms take the form
(3.9)

(3.10)

The diffusion coefficient is denoted by D. When using a reduced form of the multicomponent diffusion equation [18] derived from the complete kinetic theory to determine
the diffusion velocity components, these terms take the form
r =

1,2, ... ,N

(3.11)

DMT Ei =

phsfsUis

(3.12)

s=1

The diffusion velocity components are denoted by U.


multicomponent diffusion equation is

The reduced form of the

(3.13)

Computational Methods in Hypersonic Aerodynamics 347


and

Drs = 0.001S5SJT3[(w r + Ws)/(WrWs)]


pa;sn rs

(3.14)

Eq. (3.13) is based on the assumptions that there is no thermal diffusion and that the
same body force per unit mass is acting upon each species. X, a, and n denote the
species mole fraction, effective collision diameter, and collision integral, respectively.
Since for most turbulent mixing problems the Lewis number, which is the ratio of
the Prandtl and Schmidt (Sc) numbers, is approximately unity, the expression for the
effective diffusion coefficient is given by

pD = (Drs

;cJ

(3.15)

In Eq. (3.15), Drs can be found from Eq. (3.14) when using the multicomponent diffusion model, or from the relation (pDrs = /1-/ Se) when using Pick's law by specifying
the Schmidt number (Se = 0.22).
To calculate the required thermodynamic quantities, the specific heat for each species
is defined by a fourth-order polynomial in temperature, whose coefficients are found by
a curve fit to the available data. The molecular viscosity and the thermal conductivity
coefficients for each species are computed from Sutherland's formula. Their values for
a mixture of gases are determined from Wilke's law as follows,

(3.16)

where

[1 + (/1-s//1-r)1/2(Wr/Ws)1/4f
<Prs =

/0

vS [1

+ (ws/w r )]

1/2

(3.17)

Further details of determining the binary diffusion coefficients, Sutherland constants,


and the coefficients of the polynomials for the specific heat of each species are given
in [4, 5].
For a two-dimensional mixing flow of N species, there are (N - 1) species continuity
equations along with the global continuity equation, two momentum equations, and the
energy equation. The mass fraction of the Nth species, f N, can be found from the
following identity
(3.18)

348 Computational Methods in Hypersonic Aerodynamics


Therefore, (N + 3) coupled partial differential equations [Eq. (3.1)] need to be solved
for the vector of conserved quantities [Eq. (3.2)]. However, in an attempt to compute
the global mass conservation error, the computations are repeated by solving N species
continuity equations, that is, a total of (N + 4) coupled equations.
The explicit MacCormack [19] algorithm is used to solve the governing equations.
The present implementation of this well-documented [20, 21] predictor-corrector scheme
is based on the finite difference discretization. The type of differencing is alternated
at every other time step for symmetric computations. The stress terms [Eq. (3.4)]
are differenced in the direction opposite to those of the fluxes. The scheme is only
conditionally stable and is second-order accurate both temporally and spatially. Fourthorder damping terms are added for shock capturing.
The diffusion velocities V are calculated using two different models, which are
the complete multicomponent diffusive interaction model, and the simple binary interaction model. In the binary model, mass diffusivities of all the species are assumed
identical, and only concentration gradient effects are included [Eqs. (3.9) and (3.10)].
Whereas in the complete multicomponent model, the mass diffusivity of each species is
computed using Eq. (3.14). Then, the diffusion velocity of each species is determined
from Eq. (3.13), which requires solving (N) simultaneous algebraic equations for each
component of the velocity. It should be noted that for N species, however, the system
of N equations defined by Eq. (3.13) is not linearly independent. Therefore, one of the
equations must be replaced by the following constraint
N

LPfsV =

(3.19)

s=l

The resulting system of algebraic equations is solved using a lower-upper (LV) decomposition method. When solving (N) species continuity equations, this model [Eqs. (3.113.14)] cannot be applied, because Eqs. (3.18) and (3.19) can no longer be satisfied in
an exact manner due to the computational error.
The local mass error, LME, distribution due to the modeling of the multispecies
mixing is computed from the formula below, which is evaluated at every grid point
N

LME=p-

LPs

(3.20)

s=l

The global mass conservation error is also computed by numerically integrating the
mass along the computational domain boundaries.
The two-dimensional computational domain includes a region above the cowl where
the flow is hypersonic. The rest of the computational domain is bounded by the lower

Computational Methods in Hypersonic Aerodynamics 349


surface of the cowl and the ramp, where the supersonic flow through the internal nozzle
expands (Fig. 6). This computational domain is selected to be (18.5 h by 14 h), where h
is the throat height, and it corresponds to the longitudinal plane located at the half-span
of the internal nozzle (Fig. 3). The cowl and the ramp angles are 12 deg and 20 deg,
respectively. A fixed, boundary fitted grid is generated with appropriate clustering in
the regions where high-flow gradients are expected The global grid, which consists
of 8,839 cells, is divided into four blocks. The grid lines are contiguous across the
block interfaces, where the solutions from each side of the interface are matched. In
the normal direction, the cowl separates blocks 1 and 2, and a horizontal line extending
from the cowl tip to the downstream separates blocks 3 and 4. In the stream wise
direction, the normal line at the cowl tip separates blocks 1 and 3, and blocks 2 and
4. This multiblock approach of domain decomposition alleviates the numerical errors
that might occur if the boundaries and the interior of the cowl were included in the
computational grids [4].
The governing equations are initially solved on this fixed grid until the global error
is reduced by about 2 orders of magnitude. Then the grid is adapted to the current local
flowfield solution using the two-dimensional spring-analogy approach of [22]. This grid
adaptation procedure enhances the solution by reducing the global error by another 2
orders of magnitude.
The adaptation is done as a sequence of one-dimensional operations. For example,
the operation starts in the f. direction by redistributing the grid points according to a
specified weighting function starting from the ( = 0 line to the ( = (max line. Then
the process is repeated in the ( direction on the f. lines. The weighting function, in
the present study, is derived from the gradient of the composite function, [0.5 p + 0.3
u + 0.2 ,], a specified minimum step size, and a specified maximum step size. The (
direction adaptations are performed separately for the region above the cowl (blocks 1
and 3) and the region below the cowl (blocks 2 and 4). The f. direction adaptations are
also performed separately, first for block I, then for block 2, and finally for blocks 3 and
4 together. At the end, all these separate parts are blended together by the adaptations
applied only to the block interfaces. This practice ensures maintaining the original
shape of the cowl and the block interfaces. Further details of this flow-adaptive grid
scheme, including the necessary equations, are given in [22] and its implementation is
described in [5, 6].
4. RESULTS OF FLOWFIELD ANALYSES
The upstream conditions of the nozzle exhaust flow and the external flow are given
in Table 4.1. All of the flows are considered to be fully turbulent. Only case 1 assumes
the exhaust gases to be air. The cold gas simulating the exhaust gases is the Freon12-argon mixture for cases 2-5. Presented in Table 4.2 are the computational models

350 Computational Methods in Hypersonic Aerodynamics


for these cases. Only case 1 is computed both in two-dimensions and three-dimensions.
The computations for the other cases are in two-dimensions. Cases 1 and 5 assume
homogenous composition of the fluids everywhere and at all times. Therefore, the
species continuity equations and the terms representing the multispecies mixing are not
used for cases 1 and 5. Computed in cases 2-4 are the mixing of four species, namely,
nitrogen, oxygen, Freon-12, and argon. The diffusive mass transport model derived
from the complete kinetic theory is used in case 3, but the binary diffusion model is
used in cases 2 and 4. Only case 4 does not assume that the sum of mass ratios of
all the species is unity (Eq. 3.18).
The internal geometry for the supersonic nozzle features two corners - a lower
corner at the beginning of the ramp and an upper corner upstream of the cowl tip.
Two centered expansion fans develop around these corners. These two expansion fans
smoothly converge at the upper corner and then propagate out into the jet plume flow.
As the flow clears the internal nozzle-exit plane, two conditions can exist. When the
jet static pressure is greater than the external freestream static pressure, the jet flow
is underexpanded (Cases 1-4). When the static pressure of the jet is less than the
freestream static pressure, the jet flow is overexpanded (Case 5).
Table 4.1 Flow conditions at upstream of computational domain

Case

Flow

Fluid (by
volume)

Mach
No.

Reynolds
No. based
on (h)

Total
temp.
(OK)

Total
pressure
(kPa)

Nozzle
throat

Air:
79% N2
21% 02

1.7

192,000

475

166.0

2-5

Nozzle
throat

50%
Freon-12
50%
Argon

1.7

7500

467

172.4

1-4

External
flow

Air:
79% N2
21% 02

6.0

346,000

478

2517.0

External
flow

50%
Freon 12
50%
Argon

1.7

7500

467

172.4

~-~~~

--~~~--

Computational Methods in Hypersonic Aerodynamics 351


Table 4.2 Computational mixing models

Case
1

Multispecies
mass transE0rt
No

Yes

Yes

Yes

No

Specific heat
ratio
function of
Temperature
Temperature,
Composition
Temperature,
Composition
Temperature,
Composition
Temperature

Eqs. (3.9-3.10)

No. of partial
differential
equations
solved
4(2-0) or
5(3-0)
4+3

Eqs. (3.11-3.14)

4+3

Eqs. (3.9-3.10)

4+4

Diffusive mass
transE0rt model

4.1 Three-Dimensional Results:


The three-dimensional results obtained for Case 1 are shown in Figs. 7 through 11.
The pitot pressure contours for an 7]--<:onstant plane, located at 1.5 in from the reflection
plate, are shown in Fig. 7. Just upstream of the cowl tip, a 0.53 in thick boundary layer
is formed for the external flow. The supersonic-hypersonic mixing of air forms a shear
layer downstream of the cowl tip. This shear layer behaves like an extension of the cowl
and the flow continues to expand between the ramp and the shear layer. Two centered
expansion fans develop around the comers inside the nozzle. A small plume shock,
caused by the high-pressure expanding jet interacting with the low pressure external
flow, forms at the cowl tip and deflects downward at about -10. This shock can induce
separation in the region of the cowl tip. The extent of this separation (when it exists) is
highly dependent on the plume shock strength. The jet also expands in the stream wise
direction. The flowfield along the ramp and side ramp contains expansion waves.
The same type of flow features are present in the span wise direction and the jet
laterally expands out into the freestream. A shear layer develops between the highpressure, low-speed jet and the low-pressure, high-speed external flow. This leads to a
plume shock as the hypersonic external flow is slowed down by the expanded jet plume.
Shown in Fig. 8 are the Mach contours for a (-constant plane (approximately parallel
to the ramp at 0.3 in) of the afterbody configuration, where the nozzle, the ramp, and
the side ramp are observed. The jet expands downstream of the nozzle, and the highly
expanded lateral jet plume is clearly seen. A much thicker boundary layer develops at
the side wall in comparison to the one seen in the normal direction. The higher-pressure
jet causes the external hypersonic flow, approaching the nozzle exit plane, to experience

352 Computational Methods in Hypersonic Aerodynamics


a plume shock. As a result of the spanwise expansion of the jet and this plume shock,
a flow separation is started, and it propagates span wise along the side ramp.
A more comprehensive view of the afterbody flowfield is shown through its
crossflow Mach contours in Fig. 9. Expansion is seen in normal, spanwise, and
streamwise directions. The growths of turbulent boundary layers on the reflection plate,
ramp, and side ramp are evident. The exiting jet transforms from a rectangular shape at
the nozzle exit plane to an enlarged elliptical plume as the flow propagates downstream.
Obviously, this affects the shape of the resulting three-dimensional shear layer.
Computational off-surface pitot pressure, Pp , values are compared with the experimental results [3] in Fig. 10. Four separate rake stations (Station 1 is located at the
nozzle exit plane), each containing 25 pitot tubes, are placed at midspan on the ramp.
Here, (L) denotes the length of the rake measured approximately normal from the ramp
surface, and (s) denotes the tangential distance along the ramp surface, measured from
the 20 ramp corner. The numerical simulation of the shear layer is accomplished with
some deviations seen in the high peak values of Stations 2, 3 and 4 located at s=3.5 in,
4.54 in, and 5.54 in, respectively. According to the experimental results, a compression
wave forms at the cowl tip and extends to the third rake station at approximately -10.
The effect of this shock is seen both experimentally and computationally in the low
peak values. Computational results follow the experimental trend for the shock with
some deviation in location and strength. Station 3 reveals the largest discrepancy.
Comparisons of the computational and experimental [2] surface pressure coefficients
on the ramp and side ramp are shown in Fig. 11. Pressure values are plotted at five
spanwise locations. The first three stations (located on the ramp), which are downstream
of the nozzle exit plane, exhibit high initial Cp values before gradually declining as the
flow continues to expand down the ramp. The Cp distributions on the side ramp (1]=3.50
in and 1]=4.25 in) are relatively constant and predict slightly lower Cp values than the
final Cp value attained on the ramp. This is due to the greater spanwise expansion on the
side ramp. All of the computed Cp values compare very well with their experimental
values.
Discrepancies between the computational and the experimental results can be
attributed mainly to the grid, the turbulence model, and the uncertainties associated
with the wind tunnel data. Some improvement is possible by using a more refined
initial grid followed by adapting the grid to the flow solution as it develops. Also, the
boundary layer thickness used at the upstream of the internal nozzle flow (0.072 in)
is only assumed to be approximating the experiment, since boundary layer thicknesses
were not measured during the wind tunnel tests [2, 3].

Computational Methods in Hypersonic Aerodynamics 353


4.2 Two-Dimensional Results on Adaptive Grids
An overexpanded flow case is designed to check the solution method on adaptive
grids (case 5). The fluid is a mixture of 72% Freon-12 and 28% Argon, by mass
(46%-54%, by volume). This mixture is assumed to retain its composition unifonnly
everywhere, hence the computations are for the constant specific heat ratio of ,=1.214.
The converged solution is obtained on the second flow-adapted grid shown in Fig. 6b.
Shown in Fig. 12 are the Mach contours. The jet at the cowl tip plane is overexpanded,
which results in a shock, with a pressure ratio of 2.7, impinging on the ramp. A
separation bubble and a reflected weaker shock are observed. The interaction of the
internal and external flows is through two expansions which emanate from the cowl tip
region. The top expansion is a fan pointed upward with an included angle of about 55 .
The lower expansion is directed inward interacting with the reflected shock towards the
downstream, where the ratio of pressure to that of throat is 0.3. A small separation
zone is also detected at the upper cowl tip.
In the remainder of this section, the multispecies flow computations, i.e. cases 2-4,
will be discussed. Shown in Fig. 13 is a representative flow adaptive grid used for these
cases. Presented in Fig. 14 are the computed and experimentally measured [2] pressure
distributions on the ramp surface. All pressures are nonnalized with the pertinent
pressure values at the upstream corner. The rate of expansion of airflow (case 1) is
much higher than that of Freon-argon mixture (cases 2-4) at the corner. The difference
between the expansion rates gradually decreases, but the pressure ratios of Freon-argon
mixture are consistently higher than those of air. The computed pressure values from
cases 2-4 are very close to each other. In comparison with the experimental data,
they are initially slightly lower, then slightly higher. The computed values of surface
pressures for the air expansion (case 1) are also slightly lower than the experimental
values, but they match almost identically down the ramp.
In an attempt to assess the error associated with the assumption that the sum of
computationally obtained species mass ratios is unity (Eq. 3.18), the mass deficits of
cases 2 and 4 are computed (Fig. 15). The mass deficit is defined herein as the difference
between the numerically integrated outflux and influx of mass through the boundaries
of the computational domain. The solutions of cases 2 and 4 show convergence in
about 4000 iterations (pseudo-time steps). The mass deficit is less than 1% for case
2, where the sum of the mass ratios is assumed to be unity. This is a commonly
used assumption in similar mixing, multispecies computations, such as [20]. When
this assumption is not made, and consequently four species continuity equations are
solved for four species (case 4), the mass deficit is about 8%. A further analysis is
performed in order to find multi species mixing (Eq. 3.20) and the results are plotted in
Fig. 16. This error, of course, is identically zero for the other cases. This mass error is
occurring, as expected, within the shear layer, where the mixing takes place. It grows
in the downstream direction to a maximum of 1%.

354 Computational Methods in Hypersonic Aerodynamics


The contours of the specific heat ratio ("I) are plotted in Fig. 17. It varies throughout
the flowfield as a function of temperature and local gas composition. In case 1, however,
the variation of "I is only due to temperature. The values of "I change in the stream wise
direction and the normal direction, with the large gradients being in the shear-layer and
boundary-layer regions. Effects of the diffusive mass transport models can be observed
by comparing Figs. 17a and 17b. The model derived from the complete kinetic theory
(case 3) produces smoother shear layer and boundary layers. The binary model (cases
2 and 4) produces more oscillations, with quantitative results varying by about 2%.
These oscillations exist despite the apparent converged solution of case 2 as indicated
by an examination of Fig. 15. Therefore, the extra computational cost of the model used
in case 3 may be justified if better accuracy and oscillation-free solutions are desired.
It should be pointed out that the flows under consideration are turbulent, high-speed
flows. The differences in the results from these two models are expected to be more
pronounced for laminar, low-speed flows.
The mass fraction contours for Freon-12 and nitrogen are shown in Fig. 18. The
fluid composition at the edges of the shear layer is slightly different from its upstream
mixtures. There is a very large gradient of mixture composition through the shear
layer. Some Freon-12 and argon mixture is entrained upstream with the reversed flow
on the upper surface of the cowl. When Figs. 17 and 18 are inspected together, it is
observed that the major cause for the variation of "I in the shear layer is the change in
the composition of the multispecies fluid. In other regions, such as near the walls, the
variation of "I is primarily due to temperature gradients.

5. AERODYNAMIC DESIGN OPTIMIZATION


In the previous sections, flow analysis methods and their results were presented.
These are prerequisites to a design process, which is explained in this section. As it will
become obvious, this process is relatively more expensive in terms of computations.
Therefore, it will be demonstrated only in two-dimensions and using the inviscid flow
equations for the air flow (Euler equations). Its extension to three-dimensional, viscous,
multi species flows is straightforward, but computations are certainly more costly.
It is desired to determine the nozzle ramp shape which yields a maximum axial
thrust force coefficient, F, subject to constraints, Gj (Fig. 19). There are a number of
ways to choose the design variables [12]. Two of them are presented here: surface grid
and Bezier polynomials [23]. Since the surface grid for the inviscid analysis contains
47 points from the corner to the tip of the ramp, and the local slope at each grid rather
than its coordinates are used as the design variables, the number of design variables
(NDV) is 47 for the first choice. For the second choice, six control points are chosen
for the Bezier polynomials. Hence, NDV is 6 for this option.

Computational Methods in Hypersonic Aerodynamics 355


Mathematically, it is required to get
(5.1)

subject to
(5.2)

(5.3)
where NeON is the number of constraints, and Q is the vector of the conserved
variables of the fluid flow. XDI ower and XD upper are the lower and the upper bounds
of the design variables.
The component of the axial thrust force due to nozzle wall shape, Faxia/. is obtained
numerically by integrating the pressure, P, over the ramp and cowl surfaces. Then it
is nonnalized by the force associated with the inflow, which is parallel to the cowl
surface. For a constant inflow Mach number, Mth' the inflow force is centered at the
mid-point of the line segment kc and its value is
(5.4)

where h is the throat (th) height and 'Y is the ratio of the specific heats. By definition,
the axial thrust force coefficient, F, is given in [11] as,
F

Faxial
Fin/low

= (J~ Prampdy) + (1; Pcowldy)

(5.5)

Fin/low

This axial thrust force coefficient is subject to three physical constraints. The first
constraint requires that the static pressure at the ramp tip, PI, reaches a percentage of
the free stream static pressure, Poe, such that maximum expansion over the ramp is
reached without any reverse flow. The second and third constraints require that the
static pressure at the cowl tip, Pn , should be within specified limits of the free stream
static pressure, such that expansion waves emanating from the ramp initial expansion
do not induce any reverse flow on either the internal or external cowl surfaces.
In addition to the physical constraints stated above, there are geometrical constraints
on the configuration (Fig. 19). In order to maintain the total length of the aircraft as a
constant, the axial length of the ramp is fixed. Also, in order to maintain an acceptably
smooth aerodynamic surface shape, upper and lower limits are imposed on the local
surface coordinates, local slopes, and the Bezier control points relative to their neighbors.

356 Computational Methods in Hypersonic Aerodynamics


This nonlinear, constrained optimization problem is solved using the modified
feasible directions method developed by Vanderplaats (Ref. 24). Given a set of initial
values for the design variables, the method first drives all the design variables into a
feasible design space, i.e., the space where none of the design constraints are violated.
Then the optimum design is methodically searched for within this space. This search is
controlled by search directions which are based on the objective function gradient and,
therefore, efficiently guides the succeeding calculations towards incrementally improved
designs. Since this optimization process requires many evaluations of the objective
function and constraints before an optimum design is reached, the process can be very
expensive if a CFD analysis were performed for each evaluation. In the present study,
however, the higher fidelity flow prediction method (approximate flow analysis), which
is explained in Section 5.2, is performed during the one-dimensional searches of the
optimization process. A complete CFD analysis is performed only when new gradients
of the constraints and the objective function are needed, i.e. when the design changes
substantially. A flowchart of this overall design process is presented in Fig. 20.
5.1 Sensitivity Coefficients
The derivatives of the objective function, F, and constraints, Gj, with respect to the
design variables, XD, are defined as the sensitivity coefficients,

== d!

'fil F

dXD
'filGj

==

8! + (8~)

aXD

:~~ = :~~ + (~6)

8Q

8a:D

8Q

(5.6)

8XD
j = 1,NCON

(5.7)

These derivatives have been determined in Ref. 10 for a case with two design
variables (the inclination angles of the flat ramp and cowl surfaces to the horizontal) and
three constraint functions using two approaches; namely, the finite difference approach
and the quasi-analytical or sensitivity analysis approach. The sensitivity analysis can
be performed by one of two methods; either the direct method or the adjoint variable
method (Fig. 21). It is reported in [10] that if the number of design variables (NDV) is
greater than the number of adjoint vectors (NCON+1), the adjoint variable method is
more efficient than the direct method. Since, in the present study, the number of design
variables, XD, (47 or 6) is greater than the number of adjoint vectors (NCON + 1 = 4),
the adjoint variable method is selected to determine the sensitivity coefficients.
The governing equations for a two-dimensional, steady, compressible, inviscid flow
of an ideal gas with constant specific heat ratios written in the residual vector form are,

- ( - (- ) -)

R Q XD

8j

X D = 8~

+ 8g
8(

---------

= 0

(5.8)

Computational Methods in Hypersonic Aerodynamics 357


where j and 9 are the flux Jacobians in generalized coordinates (e, O. These equations
can be obtained by eliminating all the viscous and the mass diffusion terms in the
Navier-Stokes equations for the multispecies flow given in Section 3. Equation (5.8)
may be solved by two different methods for the analysis and the simulation of the
flowfield. The first method is the approximately factored (AF) method (Eq. 2.4), which
is explained in Section 2 and in [14-16]. The second method is using the Newton's
method [25] to solve Eq. (5.8) directly. As it will be discussed in Section 6, this direct
method is more efficient than the approximately factored method.
The sensitivity analysis approach [10] begins by differentiating Eq. (5.8) with
respect to the design variables to yield the sensitivity equation,

[~~]{a'D} = -[a~] =~

(5.9)

This equation is solved for {aQjax D }. It should be noted that Eq. (5.9) needs
to be solved for each design variable, XD; however, the coefficient matrix [aRjaQ]
needs to be factorized once and for all. The remaining partial derivatives in Eqs. (5.6)
and (5.7) can be evaluated analytically using the equations of the objective function and
the constraints. The final step is determining the values of V' F and V'Gj.
At this point, the adjoint variable method diverges from the direct method. First,
Eq. (5.9) is substituted into Eqs. (5.6) and (5.7). Then, the vectors of adjoint variables
(,\1, '\2j) are defined to satisfy the following equations,
T-

aF

J ..\1 = - -

aQ

JT,\ . _ aGi
2) -

aQ

j = 1,NCON

(5.10)

(5.11)

where JT = [aRjaQf. The adjoint variable method requires the solution of


Eqs. (5.10) and (5.11) for the adjoint variable vectors, and then, upon substitution
into Eqs. (5.6) and (5.7), yields the derivatives of F and Gj. It should be noticed that
Eqs. (5.10) and (5.11) are independent of any differentiation with respect to XD; hence,
the vectors :Xl and :X 2j remain the same for all the elements of the vectors XD.
Various methods can be used to solve the rather large sets of algebraic equations
resulting from Eqs. (5.9) or (5.10-5.11). Details and comparisons of these methods
(banded method, sparse method, iterative method, etc.) are given in [10-13].

358 Computational Methods in Hypersonic Aerodynamics


S.2 Flow Prediction Method

An approximate flow analysis method has been introduced in [10-13] to predict a


flowfield solution of a perturbed shape (Xi> + ~XD) using the flowfield solution of an
initially unperturbed shape (Xi. This method is based on a Taylor series expansion
of the vector of conserved variables Q(Xi> + ~X D) about Q(Xi and requires the
substitution of Eq. (S.9) into the Taylor expansion:
(S.12)
where
(S.13)

Equation (S.13) explicitly gives the changes in Q due to the changes in ~XD' From
here on, the term prediction is used when the flowfield is predicted using Eq. (5.13), and
the term analysis is used when the complete governing equations (Eq. S.8) are solved
with the CFD algorithms described in Sections 2 and 3.
The next logical step in the prediction method is to extend the idea of prediction
based on analysis to that of prediction based on prediction. Performing a parallel
operation to that of Eq. (S.12), we obtain a second level flowfield prediction analogous
to Eq. (5.13) by

Q(Xi> + ~xg) + ~X~)) ~ Q(Xi> + xg)) + ~Q(2)


where

Q(Xi> + ~X;(l))

is obtained by Eq. (5.13). It should be noted that

(S.14)

aQlaXD

in Eq. (S.14) is based on the predicted flowfield Q(Xi> + ~xg)). This procedure
allows flowfield solutions to be progressively "built up" from previous predictions; all
of which have the common genesis of a single initial CFD analysis solution. Thus,
a flowfield solution for a complex final shape may be obtained through incrementally
additive design variable perturbations. Otherwise, a grossly erroneous prediction is
produced if an equivalent single large perturbation is attempted.
Due to the truncation error of the first-order Taylor series, the flow prediction
is currently less accurate than the flow analysis. However, solving flowfield prediction Eq. (5.12) costs only a small fraction of solving Eq. (S.8), since (aill aQ) and
(a ill aXD) are already assembled in solving the sensitivity equation (Eq. (S.9. Therefore, for relatively small values of ~XD' significant time savings are realized at the
expense of some accuracy.

Computational Methods in Hypersonic Aerodynamics 359


5.3 Sensitivity of the Optimum Design

After the optimum design is obtained, it is desirable to determine the sensitivity


of the optimum design with respect to the problem parameters. Such information is
useful to perform trade-off analyses. For example, it may be wished to estimate what
effect a specified increase in the free stream Mach number has on the optimum thrust.
Mathematically, this requires the derivatives of the optimum values of the objective
function and the corresponding design variables with respect to the problem parameters.
The first-order sensitivity derivative method developed in [26] is adapted for the
present study. Presented in Fig. 22 is the flowchart of this process. The vector P
contains the problem parameters, which are held fixed during the optimization. Using
the superscript "op" to denote the optimum quantities, the dependence of FOP and G
on X D and P can be written as

FOP =

FOP(Q(X~(P), P),X~(P),

P)

Ga = Ga(Q(X~(P),P),X~(P),P) = 0

(5.15)

(5.16)

where Ga is a vector containing only the active constraints at the constrained maximum.
A constraint becomes an active one when its value is zero. The total optimum sensitivity
derivative of the objective function with respect to a problem parameter P is obtained
using the chain rule of differentiation. Any perturbation of the parameter P about
its value at the initial optimum must be such that the originally active constraints
remain active. For this constrained optimization problem, the first-order optimality
conditions at a local optimum (commonly known as Kuhn-Tucker conditions [27]) are
used. Combining the equations for the gradient of the objective function and the active
constraints, and using the Kuhn-Tucker conditions results in,

dFOP
dP

aFop
fJP

-raGa

+ 1/J

fJp

[(aF~p)r
fJQ

-r(a~a)rl aQ
+ 1/J fJQ
fJP

(5.17)

where 1f is a vector containing the Lagrangian multipliers. The derivatives of the


conserved variables vector, Q, with respect to the problem parameters are obtained
from the following relation,

fl(Q(XD(P),P),XD(P),P)

= 0

(5.18)

Differentiation of Eq. (5.18) with respect to the problem parameters and using Eq. (5.9)
results in,
(5.19)

360 Computational Methods in Hypersonic Aerodynamics

Solving Eq. (5.19), similar to Eq. (5.9), for (aQ lap) and using it with Eq. (5.17) yields
the sensitivities of the objective function to the problem parameters.
Alternatively, the adjoint variable method can be used when the substitution of
Eq. (5.19) into Eq. (5.17) is perfonned. Then, an adjoint vector :\, that satisfies the
following equation, is defined

JT:\ =
where J =

[aRlaQ].

[a:~p + (~~ . ~ ) ]

(5.20)

Substitution of Eq. (5.20) into Eqs. (5.17) and (5.19) gives,


(5.21)

The adjoint system of Eq. (5.20) is independent of any differentiation with respect
to the problem parameters. Also, both tenns on the right hand side are available from
the calculations of Eqs. (5.6) and (5.7). The partial derivatives, aFop lap, aCal aP and
aRlap can be evaluated analytically. Therefore, the sensitivity derivatives (Eq. (5.21
are obtained after solving Eq. (5.20), evaluating the Lagrangian multipliers, and finally
perfonning the pertinent substitutions.
Since the number of problem parameters, P, (equal to seven for the present example)
is greater than the number of the adjoint vectors, :\, (equal to one for the present
example), the adjoint variable method is more economical [10] for the present example.
6. RESULTS OF DESIGN OPTIMIZATION
Prior to discussing the actual optimization results, a series of cases (Cases 6--10) will
be presented to demonstrate the flow prediction method in Section 6.1. Subsequently, the
optimized nozzle-afterbody shape (Cases 6, 11, and 12) will be introduced in Section
6.2.
6.1 Demonstration of Flow Prediction Method
Initially, the governing equations of the flow are solved to obtain the analysis for
a flat ramp surface at 0:=10 (Case 6). Then the ramp is deflected in such a way that a
shock can be generated; a compression comer is fonned at 38% length from the ramp
from the ramp surface (Fig. 23). CFD
comer by turning the surface at an angle,
analyses are perfonned for three values of e: 2.5, 5.0, and 10 (Case 7). Finally, the
f10wfields for the above values are predicted (Case 8) using the analysis for the flat
ramp surface, that is, e=o. In other words, the =0 configuration (Case 6) is denoted by
(X D) in Eqs. 5.12-5.14, and anyone of the f:. 00 configurations (Case 8) is denoted
by (XD + 6XD)' The flow of Case 6 over the ramp is free of shocks. However, flows

e,

Computational Methods in Hypersonic Aerodynamics 361


containing compression shocks, due to ramp wall perturbation (B
based on the shock-free flow.

=fi 0), are

predicted

Table 6.1 Distinguishing Features of Flow Prediction Cases

Case

Deflection
Angle, B

Flowfield Solution Method

Analysis

2.5, So, 10

Analysis

2.So, So, 10

Prediction based
on analysis of B=O

So

Prediction based
on analysis of B=2.So

10

So

Prediction based
on prediction of B=2.5

The pressure coefficient distributions along the ramp for the Cases 6, 7, and 8
are shown in Fig. 24. The analysis (Case 7) and the predicted (Case 8) results are
indistinguishable up to the compression corner. The compression corner shocks are also
predicted very well. As expected, discrepancies begin to appear for the larger B values,
i.e. larger design variable changes. Notice that a discontinuous physical phenomenon
(shock) is predicted based on a flow which does not have that phenomenon (shockfree). The maximum deviation is only 2% for B=2.So, but it increases to 22% for
B=lO. These deviations can be attributed to the nonlinear nature of the shock. This is
a typical trend when a nonlinear problem is solved using a method which includes some
kind of local linearization. Therefore, it can be positively concluded that the prediction
method, due to truncation error, exhibits increasing inaccuracies as the deflection size,
6XD, increases and eventually produces unacceptable solutions when the deflection
becomes too large.
An issue which appeals to the curiosity is the success of the present prediction
method for the off-surface flow, particularly when an existing change in a configuration
is enlarged; for example, predicting the flow for B=So when the flow of B=2.So is given.
Shown in Fig. 2S are three sets of density contourS in comparison: the flow analysis of
B=2S, the flow predictions of B=S.O based on first the flow analysis of B=2.So (Case
9) and then the flow predictions of B=2.5 (Case 10). Two points are noteworthy here.
First, Case 9 aims at predicting the flow due to an enlarged change (B from 2S to
S.OO), but not predicting a new physical phenomenon; that is, the prediction method
is given a shock with which it begins. The comparison of this case with the analysis
is satisfactory, despite the fact that the shock is rather strong. Secondly, this figure

362 Computational Methods in Hypersonic Aerodynamics


illustrates the feasibility and quality of the second level prediction of Eq. 5.14. For
example, the prediction of ()=5 flowfield (Case 10) is based on the Case 8 prediction
for ()=2.5, which is based on Case 6 analysis. Since the truncation error occurs twice
and progressively during this process, although associated with smaller perturbations,
the agreement of Case 10 is slightly less successful than that of Case 9. Therefore,
it may be concluded that higher level predictions may be further used for coarse but
efficient estimates of highly perturbed shapes.
6.2 Shape Optimization of Nozzle-Afterbody
Three different and arbitrary shapes are chosen as the initial design shapes for the
ramp; namely, a flat ramp surface at 0:'=10 (Case 6), a concave surface with its axis
(the straight line connecting the corner point and the ramp endpoint) at 0:'=25.7 (Case
11), and a convex surface with its axis at 0:'=29.5 (Case 12). The slope of initial
ramp expansion is 35.0 for the concave shape and 23.5 for the convex shape. The
reason for starting the optimization from three different initial shapes is to determine
how close the resulting optimized ramp shapes are to each other. Ideally, they should
be identical irrespective of their initial shape, so that the designer using this method
in the production mode can start with any shape that is convenient. Shown in Fig.
26 is the comparison of the optimum shapes of Cases 6, 11, and 12 along with their
initial shapes. The optimum shapes are almost identical for 70% of the surface and
show a small difference towards the tip. When the axis angles, 0:', of the optimum
shapes are compared, it can be seen that the difference between Cases 11 and 12 is
indistinguishable (less than 0.3%) and that of Case 6 differs from them by only 3%.
The effect of the shape optimization on the interior flowfield is just as pronounced
as it is on the surface properties. The Mach number contours of both initial and
optimum configurations of Case 6 are presented in Fig. 27. The expansion patterns are
significantly different. The rate of expansion is much higher inside the nozzle for the
optimum shape, which results in a higher Mach number and less underexpanded jet at
the nozzle exit plane. The consequence of this is evidenced in the shear layer, which
is thinner and has a smaller angle with the horizontal for the optimum shape. Also, the
expansion along the external part of the nozzle ramp is no longer predominantly in the
streamwise direction, but a significant portion is in the normal direction. This indicates
that cancellation of the cowl corner centered-expansion waves occurs at the optimized
ramp surface, which is a characteristic feature of bell-type nozzles.
Plotted in Fig. 28 are the histories of the objective functions, F, during the
optimization iterations (or commonly known as levels). The initial F value of Case
11 is the highest, and all three shapes converge to an optimum F value within 14
optimization iterations. Cases 6 and 12 have identical optimum F values to the fourth
significant digit, and that of Case 11 differs from them by less than 0.5%.

-'

---------

Computational Methods in Hypersonic Aerodynamics 363


The computational time for each one of the shape optimization cases is about 3.5
hours on the CRAY-YMP of Numerical Aerodynamic Simulation (NAS) of NASA.
For example, Case 6 requires 180 evaluations of the objective function over the course
of 14 optimization iterations. At the end of each iteration, there is a CFD analysis
accompanied with the objective function evaluation for the new improved shape. This
means that CFD analysis is performed 14 times, whereas the flow prediction calculation
is performed 166 times. Comparing the computational times of an analysis (rv 175
seconds) and a prediction (rv 18 seconds), it can easily be realized that the aerodynamic
optimization procedure is more efficient by employing the present prediction method.
To demonstrate the relative efficiencies of the present design optimization methods,
Case 11 is solved using five different procedures and their results are given in Table 6.2.
In Procedure 1, the sensitivity coefficients, VF and VG, are computed by the traditional
finite-difference method. In contrast, they are computed by the sensitivity analysis (SA)
approach in the rest of the procedures, i.e. Procedures 2-5. The design variables in
Procedures 1-3 are the local slopes at each surface grid point (47 of them), but they
are the Bezier control points (6 of them) in Procedures 4 and 5. The third important
difference between the procedures is the method to solve the flow equations; all but
the last procedure employ the traditional approximate factorization (AF) method and
the last procedure solves these equations directly using the Newton's method. Finally,
Procedures 3, 4, and 5 employ the flow prediction method, but Procedures 1 and 2 use
the complete CFD analyses only.
The most efficient method is Procedure 5 and it requires 2.2 hours of computing on
the CRAY-2 computer of NASA Langley Research Center. The most time is required
by Procedure 1, which uses all the traditional methods. The memory requirement
is of the same order of magnitude for all procedures. Part of the reasons for such
disparity in computational efficiencies may easily be understood by inspecting the last
two columns of Table 6.2. It is noteworthy to point out that Procedure 2 requires less
CFD analyses than Procedure 1. This is due to the values of the sensitivity coefficients,
which are obtained much more accurately when the sensitivity analysis approach is
used. Consequently, the optimizer provides a converged solution much quicker. Further
details of such efficiency considerations are given in [10-13].

364

Computational Methods in Hypersonic Aerodynamics

Table 6.2 Normalized Computational Times Required


by Different Aerodynamic Optimization Procedures

Normalized
Execution
Time

Required
Memorey [MW]8

NO.ofCFD
Analyses

No. of
Approx. How
Analyses

1) AF+slopes+FD

49.2

6.1

595

2) AF+slopes+SA

19.7

6.1

236

3) AF+slopes+SA

3.75

6.1

230

4) AF+Bezier+SA

2.04

5.4

15

172

5) Newton's+Bezier+SA

l.00b

5.4

15

172

Optimization Procedure

AF: Approximate Factorization


FD: Finite Difference for sensitivity
SA: Sensitivity Analysis for sensitivity coefficients
a grid size (41x53)
b cpu time = 2.2 hrs on Cray2

7. REFERENCES
1. Oman, R. A., Foreman, K. M., Leng, J., and Hopkins, H. B. Simulation of
Hypersonic Scramjet Exhaust. NASA-CR-2494, 1975. Validation of Scramjet
Exhaust Simulation Technique at Mach 6. NASA-CR-3003, March 1979.

2. Cubbage, J. M. and Monta, W. J. Surface Pressure Data on a Scramjet Nozzle


Model at a Mach 6 Using a Simulant Gas for the Engine Exhaust Flow. NASP
Contractor Rept. 1058, NASP-JPO, WPAFB, Ohio, Aug. 1989.
3. Monta, W. J. and Cubbage, J. M. Pitot Rake Surveys for a Scramjet External Nozzle
Model at Mach 6 Using a Simulant Gas for the Engine Exhaust Flow. Prospective
NASA Technical Report, NASA Langley Research Center, Hampton, VA, 1991.
4. Baysal, 0., Engelund, W. C., and Tatum, K. E. Navier-Stokes Calculations
of Scramjet-Afterbody Flowfields. Advances and Applications in Computational
Fluid Dynamics. (Ed. Baysal, 0.), FED-Vol. 66, pp. 49-60, ASME, New
York,Nov. 1988.
5. Baysal, 0., Engelund, W. C., Eleshaky, M. E., and Pittman, J. L. Adaptive Computations of Multispecies Mixing Between Scramjet Nozzle Flows and Hypersonic
Freestream. AIAA Paper 89-0009, Jan. 1989.

Computational Methods in Hypersonic Aerodynamics 365


6. Baysal, 0., Eleshaky, M. E., and Engelund, W. C. Computations of Multispecies
Mixing Between a Scramjet Nozzle Flow and Hypersonic Freestream. (AIAA)
Journal of Propulsion and Power, Vol. 8, No.1, Jan-Feb. 1992.
7. Baysal, 0., Eleshaky, M. E., and Engelund, W. C. 2-D and 3-D Mixing Flow
Analyses of a Scramjet-Afterbody Configuration. Paper No. 14, Proceedings of
International Conference on Hypersonic Aerodynamics, The Royal Aeronautical
Society, Sep. 1989.
8. Baysal, O. and Hoffman, W. B. Computation of Supersonic-Hypersonic Flow
Through a Single-Module Scramjet Nozzle. Proceedings ofIMACS 1st International
Conference on Computational Physics, pp. 52-59. University of Colorado, Boulder,
Co., June 1990.
9. Baysal, O. and Hoffman, W. B. Simulation of 3--D Shear Flow Around a NozzleAfterbody at High Speeds. Advances in Numerical Simulation of Turbulent Flows,
(Ed. I. Celik, et al.), FED-Vol. 117, pp. 75-82, Joint Meeting of ASME/JSME, June
1991. Also, (ASME) Journal of Fluids Engineering, Vol. 114, No.1, Mareh 1992.
10. Baysal, O. and Eleshaky, M. E. Aerodynamic Sensitivity Analysis Methods For
the Compressible Euler Equations. Recent Advances and Applications in CFD,
(Ed. Baysal, 0.), FED Vol. 103, pp. 191-202, AS ME-Winter Annual Meeting,
Nov. 1990. Also, (AS ME) Journal of Fluids Engineering, Vol. 113, No.4,
Dec. 1991.
11. Baysal, O. and Eleshaky, M. E. Aerodynamic Design Optimization Using Sensitivity
Analysis and Computational Fluid Dynamics. AIAA-91-0471, Jan. 1991. Also,
AIAA Journal, Vol. 30, No.2, Feb. 1992.
12. Baysal, 0., Eleshaky, M. E., and Burgreen, G. W. Aerodynamic Shape Optimization Using Sensitivity Analysis on Third-Order Euler Equations. AIAA Paper
No. 91-1577 CP, Proceedings of 10th Computational Fluid Dynamic Conference,
pp. 573-583, June 1991.
13. Eleshaky, M. E. and Baysal, O. Airfoil Shape Optimization Using Sensitivity Analysis on Viscous Flow Equations. Multidisciplinary Applications ofCFD (Ed. Baysal,
0.), ASME-FED, Vol. 119, pp. 26-36, Winter Annual Meeting, Dec. 1991.
14. Baysal, 0., Fouladi, K., and Lessard, V. R. Multigrid and Upwind Viscous
Flow Solver on 3-D OverlappedlEmbedded Grids. AIAA Journal, Vol. 29, No.6,
pp. 903--910, June 1991
15. Baysal, 0., Fouladi, K., Leung, R. W., and Sheftie, J. S. Interference Flows Past
Cylinder-Fin-Sting-Cavity Assemblies. AIAA Paper No. 90-3095CP, Proceedings
of 8th Applied Aerodynamics Conference, pp. 884-892. Also, Journal of Aircraft,
Vol. 29, No.1, Jan. 1992.

366 Computational Methods in Hypersonic Aerodynamics


16. Anderson, W. K., Thomas, J. L., and Van Leer B. Comparison of Finite Volume
Flux Vector Splittings for the Euler Equations. AIAA Journal, Vol. 24, No.9,
pp. 1453-1460, Jan. 1986.
17. Baldwin, B. and Lomax, H. Thin Layer Approximation and Algebraic Model for
Separated Flows. AIAA Paper 78-257, Jan. 1978.
18. Williams, F. A. Combustion Theory, 2nd ed., pp. 628-649, Benjamin-Cummings
Publishing, Menlo Park, CA, 1985.
19. MacCormack, R. W. The Effect of Viscosity in Hypervelocity Impact Cratering.
AIAA Paper 69-354, April 1969.
20. Drummond, J. P., Rogers, R. C., and Hussaini, M. Y. A Detailed Numerical Model
of a Supersonic Reacting Mixing Layer. AIAA Paper 86-1424, June 1986.
21. Baysal, O. Supercomputing of Supersonic Flows Using Upwind Relaxation and
MacCormack Schemes. (ASME) Journal of Fluids Engineering, Vol. 110, No.1,
pp. 62-68, March 1988.
22. Nakahashi, K. and Deiwert, G. S. Adaptive Grid Method with Application to Airfoil
Flow. AIAA Journal, Vol. 25, No.4, pp. 513-520, April 1987.
23. Faux, I. D. and Pratt, M. G. Computational Geometry for Design and Manufacture.
Ellis Horwood Ltd. Publishers, West Sussex, England, 1979.
24. Vanderplaats, G. N., An Efficient Feasible Direction Algorithm for Design Synthesis
AIAA Journal, Vol. 22, No. 10, pp. 1633-1640, October 1984.
25. Venkatakrishnan, V. Newton Solution of Inviscid and Viscous Problems. AlAA
Journal, Vol. 27, No.7, pp. 885-891, July 1989.
26. Sobieski, J. S. and Barthelemy, J. M. Optimum Sensitivity Derivatives of Objective
Functions in Nonlinear Programming. AIAA Journal, Vol. 21, No.6, pp. 913-915,
June 1983.
27. Reklaitis, G. V., Ravindran, A., and RagsdeU, K. M. Engineering Optimization
Methods and Applications. 1st ed., pp. 191-203, John Wiley and Sons, New York,
1983.

8. ACKNOWLEDGEMENT
This chapter was extracted from References [4-12], which were written by the
author and his valuable graduate students (G. W. Burgreen, M. E. Eleshaky, W. C.
Engelund, W. B. Hoffman). NASA Langley Research Center provided support under
grants NAG-1-811 and NAG-1-1188. David S. Miller was the technical monitor.

----------

Computational Methods in Hypersonic Aerodynamics 367

I
I
I
I

Plume boundary :

Bow shock

I
I
I
I
I

Cowl

~------------------ ____ J

Inlet shock

Region of interest
in the present study

Figure 1 A schematic of the flowfield around a generic hypersonic vehicle.

16.73h

Argon-Freon resevoc

(a)
h=O.6 in.

(b)
.Flow fence for quasi 2-D flow

Figure 2

A wind tunnel model of a single-module scramjet nozzle and afterbody: (a)


sectional view, (b) isometric view.

368 Computational Methods in Hypersonic Aerodynamics

41x65x41
65x65x41
65x65x41
33x33x41
33x33x41
65x33x41
65x33x41

Figure 3

grid 1
grid 2
grid 3
grid 4
grid 5
grid 6
grid 7
grid 8

Three-dimensional grid for the computational model of the nozzle-afterbody.


The grid is made of eight blocks and 808,848 cells.

Computational Methods in Hypersonic Aerodynamics 369

8h65x41

Figure 4

GRID I

Mach number contours of the flow inside the rectangular duct preceding the
internal nozzle (Case 1).

(M~)ext
97x65x4]
97.-25.4]
97x25x4]

=6.0
GRID]
GRID 2
GRID 3

Figure 5 Mach number contours of the external flow past the double comer preceding
the cowl and the external nozzle (Case 1).

370 Computational Methods in Hypersonic Aerodynamics

AFTER-BODY GRID

(a)

IDIMXJDIM=

14.01.21

12.000

6.000

4 000

.000 +----.----,----.-----,----,----,---,----,--~-___,
20.000
12 000
16.0ro
B.OOO
4.000
.000

(b)

Figure 6

Grids for two-dimensional computations with 20,305 cells: (a) fixed grid,
(b) flow adapted grid for Case 5.

Computational Methods in Hypersonic Aerodynamics 371

10.9

4.7

(
3.2

Figure 7

Pitot pressure contours in the internal nozzle and ramp region for an
constant plane (Case 1).

1]-

---~~~
---c~,...------6.1

5.6

4.7__
1.9

lr\
2.8

Figure 8

Mach contours for a (-constant plane of the nozzle-afterbody (Case 1).

372 Computational Methods in Hypersonic Aerodynamics

Figure 9

Mach contours depicting various crossflow (~-constant) planes of the nozzleafterbody (Case 1).

Computational Methods in Hypersonic Aerodynamics 373

Computation

Experiment
1:;.

----------

17.5

5
5
5
5

.........................

-------

Station
= 2.18 in.
= 3.54 in.
= 4.54 in.
= 5.54 in.

15.0
12.5

Pp
(psia)

2.5

o
Figure 10

1.0

2.0

3.0

4.0

5.0

L (in.)

Comparisons of computational and experimental [3] off-surface pitot pressure (Case 1).

1.00
Computation Experiment
o

0.75

1:;.

o
C p 0.50

"V

Station

=0.375 in.
T) =2.250 in.
T) =2.625 in.
T) =3.50 in.
T) =4.25 in.
T)

2.5
5.0
7.5
Surface length,s (in.)
Figure 11

10.0

Comparisons of computational and experimental [2] surface pressure coefficients (Case 1).

374 Computational Methods in Hypersonic Aerodynamics

(a)

1.7

MRCH

CONTOURS

MIN VRLUE=

0.0000

MqX=

2.%25

(b)

1.7

1.7

_ _- - - - - 2 . 3 -

MRCH

Figure 12

CONTOURS

MIN VRLUE=

0.0000

MRX=

2.9733

Mach number contours of Case 5 where the flow is overexpanded: (a)


solution on the fixed grid shown in Fig. 6a, (b) solution on the adapted grid
shown in Fig. 6b.

Computational Methods in Hypersonic Aerodynamics 375

12.000

III

a.ooo

4.000

.000

+-----,---.,..------,----.------r--.,------r---,------,r~-.,

.000

Figure 13

4.000

a.ooo

12 000

20.000

16.000

A flow adapted grid for a flow which is underexpanded (Cases 1-4).

1.0
Case

.8

Exp.

CFD

1
P
P3

.6
.4

-----

.............

---

.2

12

16

20

Normalized distance, X/h

Figure 14

Comparisons of computed (2-D) and experimental [2] surface pressures.

376 Computational Methods in Hypersonic Aerodynamics


40
Case

30

.~

';

--

20

0
CII
CII

co

,,

10

,,
"-

1000

2000

...

3000

--

4000

5000

No. of Iteration

Figure 15

Global mass conservation error for Cases 2 and 4.

LOCAL MASS ERROR CONTOURS

MAX: 0.01 MIN:0.001 DEL: 0.001

CASE 4

Figure 16

Local mass error due to mixing of species (Case 4).

Computational Methods in Hypersonic Aerodynamics 377

(a)
GAMMA CONTOURS

MAX: 1.40 MIN: 1.18 DEL: 0.015

CASE 2

~1.400

GAMMA CONTOURS

MAX: l.4D MIN: 1.18 DEL: 0.01

CASE 3

Figure 17

Contours of specific heat ratio: (a) Case 2, (b) Case 3.

378 Computational Methods in Hypersonic Aerodynamics

(a)
FR12 MASS FRACTION CONTOURS
MAX

= 0.751

-...

MIN

= 0.1

DEL = 0.1

.766\

p;r.-""

"'1.;;,rre~~*~";:j-;"~~~r!::.::':

(b)
N2 MASS FRACTION CONTOURS
MAX = 0.766 MIN = 0.1 DEL = 0.1
CASE 3

Figure 18

Mass fraction contours of Case 3: (a) Freon-12, (b) Nitrogen.

M~=

Y'cowl

6.0
~

P=12

Rx 'cowl ~

~ m).c;q~i}~0//~,

Options for Design Variables


47 Surface Slopes
ill

PO,th

M,h = 1.665

6 Bezier Control Points 0

li.l .th

Reservoir

g'
e:..
~

Y'ramp

I"""

Ramp Axial Length = 17h

,/

S'

~
~
::s

(')'

Figure 19

Physical and geometrical fonnulation of the nozzle-afterbody optimization


problem.

)(1)

a
~

~.

VJ

380 Computational Methods in Hypersonic Aerodynamics

Initial Design Variables

XD

Problem Parameters

Evaluate
F G J's

YES

NewXn
( = initial XD in first pass )

Approximate Flow
Analysis

Computing the
Sensitivity of
the Optimum
Design w.r.t.P

YES

Flow Analysis (CFD)


(VUMXZ3)
Computing Sensitivity
Coerticients(Gradients)
V F V G'j
using Sensitivity AnaJysis
Approach

C:==STOP=::>~~--------------~

Figure 20

Flowchart of the aerodynamic shape optimization method.

I . DIRECT METHOD

II. ADJOINT VARIABLE METHOD

ali ~ ~ ali
aij 'aij' aij 'axo

f
r:;.

g'
e:..

Solve

(ali/aij)T A2

)=

(aGJ/aij)

0..
Vl

5'

~
~

n'

t
~.

Figure 21

Flowchart of the sensitivity analysis (quasi-analytical) methods (preoptimization sensitivides).

00
......

382 Computational Methods in Hypersonic Aerodynamics

Assemble
ali a FOP aeu ali a FOP aeu
-=, ---=-, -----=- , --== , --- , --=-,
ap
ap aQ aQ
ap
aQ
ali a FOP aG a

-=-- , ---=---- , ----::::-aXD

aXD

aXD

Factorize ( iJ Ii / a Q )T

~
Compute sensitivity coefficients
VFop , Ve a
using sensitivity analysis approarch

Compute the Lagrangian Multipliers ,\j!


\j!=_[(vea)T(vea)]l (vea)T V FOP

~
Solve
iJR)T - _(iJFUP ae a
A - --=-+--=( aQ
iJQ
aQ

-=

-)

\jI

Compute

dF~ =iJF~ +\j!Ta~_iTa~


dP

Figure 22

iJp

ap

dP

Flowchart of the adjoint variable method to determine the sensitivity derivatives (post-optimization sensitivities).

Computational Methods in Hypersonic Aerodynamics 383

At~3

Deflection
--..:.0 Angle,

~i
-----

------

x/h

= 0.0

Figure 23

-- -- -- -.

x/h"; 6.5

x/h = 17.0

Description of the flowfield prediction problems.

5
Corresponding

Analysis

Dl:OCClion
Angle,

Prediction

N/A

0.0'
2.5 0
5.0 0
10.00

0
0

b.

11

13

15

17

Normalized Axial Distance, x/h

Figure 24

Surface pressure coefficient distributions along the ramp for various deflection angles (Case 6-8).

384 Computational Methods in Hypersonic Aerodynamics


Analysis
Prediction based on Analysis of e =2S
Prediction based on Prediction of e = 2.5

Figure 25

Comparisons of density contours for () = 5 ramp deflection (Cases 7, 9


and 10).

::::::::=.r:

-->-

...oJ

I:

.;::
II>

-2

IklTI~L

is
-;
S

..

FLH

SHAPf _

- --=-

-4

---.;;:

Z
"tl

...

OJ

-;
S

..

-6

IklrlAL

'-VI

COI/CAYE:

tl-f<

co",

Y~-t

.,.,'~

-8

-10
-2

18

Normalized Axial Distance,

Figure 26

SH~pE:

X/h

Comparisons of final optimized ramp shapes (Case 6, 11, and 12).

Computational Methods in Hypersonic Aerodynamics 385

Axial Thrust Coefficient = 0.1157

(a)

0./~

Flat Initial Ramp Shape

Axial Thrust Coefficient = 0.1524

'.038-

36~

(b)

Figure 27

Optimized Ramp Shape

(a) Mach contours of the initial ramp shape (Case 6).


(b) Mach contours of the optimized ramp shape (Case 6).

386 Computational Methods in Hypersonic Aerodynamics

0.16

0.15

tJ.,

0.14

'"
2

F:

0;

.;:;:

Initial
Ralnp

0.13

Codf. of Axial Thrust, F


Symbol

Shape

Initiol Design Optimum Design

<t:

Flat

---0-

0.1157

0.1524

'0

Concave

--0--

0.1416

0.1517

Convex

----lr-

0.1029

0.1524

0.12

'u'"
S

'"0

Fgain: 8 - 50 %

14CFD
-166 Flow Predictions
CRAY-YMP: -3.5 hr8,

0.11
!iF < 0.5 %
0.10
0

10

12

14

Number of Optimizer Iterations

Figure 28

Optimization history of the objective function for Cases 6, 11, and 12.

Chapter 11:
The Computation over Unstructured Grids of
Inviscid Hypersonic Reactive Flow by Upwind
Finite-Volume Schemes
J.-A. Desideri
Institut National de Recherche en Informatique et en
Automatique, Centre de Sophia Antipolis, 2004 Route des
Lucioles, B.P. 109, 06561 Valbonne Cedex, France
ABSTRACT
Finite-volume and finite-element solutions to the Euler equations are presented for the
computation of inviscid hypersonic reactive flows around blunt bodies. For steadystate efficiency, the recommended method is an implicit (pseudo-)time-integration algorithm employing local timestepping. Robust spatial approximations are constructed
from either first-order or M. U .S.C.L.-type quasi second-order upwind schemes in which
the flux-vector splitting accounts for a mixture of 5 perfect gases (N2' O 2 , NO, N
and 0) in the equation of state and for a local value of the parameter /, here defined as the ratio of enthalpy to internal energy. Equilibrium flow solutions are found
by combining the basic method with a classical algebraic model, solved by Newton's
method. Solutions to recently proposed non-equilibrium chemistry models are found
by integrating forward in time additional species convection equations with source
terms. In such "finite-rate chemistry" equations, the convection terms are split in a
manner modeled on the treatment of the analogous term in the continuity equation.
The source terms are also implicited and large timesteps can be employed. The effects on the equilibrium/non-equilibrium flow of the freest ream Mach number, Moo,
and the nose radius of curvature, Rb, are evaluated. In particular, it is verified that
as Rb ~ 00 (Moo being fixed), the non-equilibrium flow solution slowly approaches
the equilibrium solution. Finally, a linear theory has revealed that the inviscid nonequilibrium blunt-body flow is singular at the stagnation point. As a result, for such
flow, grid convergence cannot be achieved unless extremely small grid spacing is employed near the wall. However, an alternative is to apply a post-processing procedure
which enforces equilibrium conditions at the stagnation point, to the standard numerical solution obtained on a grid of reasonable size; this permits the theoretical
solution (apart from a small truncation error) to be attained cheaply.

388 Computational Methods in Hypersonic Aerodynamics


INTRODUCTION
This article presents numerical methods to include chemical dissociation effects in the
computation of the hypersonic inviscid part of the flow around a typical blunt body.
In the boundary layer, chemical dissociation, and more generally, departures from
the perfect gas model (viscosity, rarefaction, catalycity, etc.) play important roles, in
particular in the computation of the wall temperature. These phenomena are beyond
the scope of the present study, but should be included in a global calculation, or in a
sophisticated boundary-layer correction calculation. Concentrating on the main flow
outside the boundary layer, the solution to the Euler equations is assumed to contain
the essential features. This study is motivated by the need to simulate the critical
phase in the atmospheric reentry of a space vehicle such as the European shuttle
Hermes, which occurs at altitudes around 60 km or higher and at a hypersonic flight
Mach number

(1)

(Voo, coo: freestream velocity and sound speed) that may range from 15 to 25. As
a result, a strong bow shock in front of the vehicle, as depicted in Figure 1, causes
important variations of the physical values of the air flow. Upstream of the shock, the
flow is uniform, the temperature is low and air is a mixture of uniform composition,
of which the main compounds are nitrogen (79%) and oxygen (21%). Through the
shock, the temperature increases extensively. For a calorically perfect gas in steady
flow, the static temperature T8 is constant along a streamline and given by
T. = 1 +

Too

'

-1 M2
2
00

(2)

where, is the ratio of specific heats (r = 1.4 for air). This temperature is also the
temperature at the stagnation point. For example, with Too = 231 K and Moo = 25
we obtain T8 '" 29000 K. Then obviously, the hypothesis of inert gas is completely
inaccurate. Long before this temperature T. is reached, the air molecules dissociate
causing an important fall of the temperature, since the dissociation reactions are
endothermic. For these reasons, the simulation of hypersonic flow to be pertinent, in
particular with respect to the temperature field computation - a crucial issue to the
designer - must involve some evaluation of the chemical dissociation phenomena [2].
In order to incorporate chemical dissociation effects in a flow solver, a first approach can be realized by coupling the Euler equations with a simple algebraic
equilibrium-chemistry model. This approach has the merit to indicate the trends
from a perfect (inert) gas to a dissociating gas model. However, equilibrium is not
realistic at high Mach numbers (Moo = 25), for the typical characteristic lengths cales
associated with the chemical relaxation phenomena are comparable with the typical
geometrical characteristic length of space vehicles (both between 0.1 m and 1 m). It
is then necessary to evaluate the kinetics of chemistry via a non-equilibrium model.
Furthermore, at high temperatures ionization may also take place.
Flow computations by modern methods incorporating real gas effects appeared in
the early 1970's with the development of the American space shuttle. Among the ear-

Computational Methods in Hypersonic Aerodynamics 389


liest numerical works, we may cite an article by C. P. Li [3] in which the compressible
Navier-Stokes equations were solved at low Reynolds numbers, past a sphere, by the
MacCormack explicit predictor/corrector method [4]. The employed non-equilibrium
model accounted for 6 species, including NO+ (and electrons). In 1975, Rakich et
al. [5] applied a method of characteristics to solve the supersonic portion of 2-D and
3-D inviscid non-equilibrium flows. The 5-species, 18-reaction chemical model they
described has since been employed by many other authors. In particular, in 1976,
Rizzi and Bailey [6] employed a space-marching finite-volume method to solve the
supersonic reactive flow past a 3-D blunt body. In 1981, Ramshaw and Cloutman [7]
focussed on the numerical treatment of a complex reactional scheme of species convection equations in which both equilibrium and non-equilibrium reactions are present.
Portions related to the kinetics have been numerically treated explicitly, whereas equilibrium reactions implicitly. In 1984, Prabhu and Tannehill [8] solved the parabolized
Navier-Stokes equations around the space shuttle Orbiter by a space-marching implicit factored finite-difference method, for an ideal gas (-y = 1.2), and for equilibrium
aIr.
A detailed description of a complex chemical model including 11 species and
accounting for dissociation, ionization and vibrational excitation has been given by
Park [9], who computed the non-equilibrium flow in a duct by a centered, factored,
implicit finite-difference method.
Collela and Glaz [10] have been, to our knowledge, the first to consider the exact
solution to the Riemann problem for gas dynamics with a general convex equation
of state. They presented efficient and accurate calculations of the 2-D reflection of a
planar shock by an oblique surface. With the same hypothesis, Glaister [11] developed
an extension to Roe's Approximate Riemann Solver and applied it to the 1-D shockreflection test problem in the cases of three different equations of state. Eberhardt and
Palmer [12] in 1986 developed an upwind TVD scheme (in an implicit approximatefactorization time-integration algorithm) to compute 2-D or axisymmetric inviscid
equilibrium flows. The modified upwinding scheme is based on pressure gradients
that account for the equation of state of the dissociated gas. Drummond et al. [13]
employed a spectral approximation combined with a Runge-Kutta time-integration
method to solve quasi-one-dimensional chemically reacting inviscid flowfields at low
supersonic regime. In [14], J-L. Montagne proposes a flux-vector splitting applicable
to an inviscid flow at chemical equilibrium, for a general equation of state, and presents
numerical results for the shock-tube problem and a blunt-body problem. Montagne
et al. [15] extended Roe's approximate Riemann solver, and the Steger-Warming and
the van Leer flux-vector splittings to the case of a general equation of state. They presented a variety of solutions to the shock-tube problem. They implemented and tested
a second-order symmetric TVD explicit scheme and a second-order MUSCL-type upwind scheme based on their theoretical results. They concluded that all approaches
provide comparable accurate solutions. In [16], implicit and 2-D extensions are described and applied to complex flow calculations. In a recent publication, Candler
and MacCormack [17] presented a method to solve the Navier-Stokes augmented to
include an equation for each of five gas species and three vibrational energies. Their
method applies flux-vector splitting on the entire set of partial-differential equations.
An efficient implicit solution procedure is based on tridiagonal inversion in the direc-

390 Computational Methods in Hypersonic Aerodynamics


tion normal to the body, and Gauss-Seidel line relaxation in the axial direction. They
made computations of a sphere-cone problem at high Mach numbers.
Finally, it should be noted that hypersonic flows can also be computed by shockfitting. Botta et al. [18] propose to compute non-equilibrium reacting flows about
blunt bodies by the "A-scheme" in which the known I-D solution behind the shock is
employed. They also present an interesting discussion on the effect of the Damkholer
number on the flowfield.
This chapter presents first a basic finite-volume/finite-element methodology to
construct a robust solver for the Euler equations applicable to hypersonic flow computations. This approach is first described in the next section in the case of an ideal
gas. It is based on flux-vector splitting and explicit timestepping. Implicit timestepping and quasi-second-order formulations are constructed subsequently. The method
is then extended to compute flows at chemical equilibrium, and a numerical experimentation is conducted. In the last section, a solver for non-equilibrium chemistry
is presented, and the effects on the flow of freest ream Mach number and nose radius of curvature are evaluated; comparisons are made with the equilibrium solutions.
A theoretical analysis of the steady inviscid non-equilibrium flow in the vicinity of a
stagnation point is made, and its impact on the validation of numerical methodologies
for inviscid non-equilibrium flow is examined.
Most of the material related to the first-order explicit algorithms and the associated numerical experimentation have already been presented in [1].
BASIC SOLVER FOR THE EULER EQUATIONS
In this section, the basic method employed to solve the Euler equations (in the case
of an inert gas) is briefly described. The essential ingredients are a finite-volume formulation applicable to a structured or unstructured domain discretization, upwinding by flux-vector splitting, and the van Leer M.U.S.C.L. approach [19] for spatially
second-order extensions. For more complete descriptions, the reader is referred to
the numerous publications in which these techniques were originally developed in the
finite-element context [20, 21]. In the subsequent sections, the method is generalized
to include the chemical dissociation effects.
Governing equations and discretization
In order to capture conservative solutions to the steady Euler equations in a mixed
subsonic-supersonic region of the hypersonic flow over a two-dimensional blunt body,
the time-dependent Euler equations are written in conservation-law form,

(3)
in which

(:u)
pv

'

G(W)

=(

::v+ )

pv

(4)

v(E + p)

where p is the density, u and v are the x and y velocity components, E is the total

Computational Methods in Hypersonic Aerodynamics 391


energy per unit volume and p is the pressure. For a perfect gas,

(5)
in which '"Y is the ratio of specific heats, that is, '"Y = 7/5 for a (mixture of) diatomic
gas(es). Note that this equation does not hold for air when chemical dissociation
occurs.
To prepare the discretization, the Euler equations are cast in integral form,

VA,

~tJJ.A Wdxdy +

iA

(F(W)nx

+ G(W)n y ) ds = 0

(6)

in which A is an arbitrary control area, 8A its oriented boundary, and

(7)
the outward normal.
A typical domain of integration is shown on Figure 2. It is bounded by an inflow
boundary rin along which the supersonic flow is given, a supersonic outflow boundary
rout along which no specifications are made, and the body surface rb defined by a shape
function I and a geometricallengthscale Rb (usually the nose radius of curvature), so
that the body surface equation is

(8)
in which { and." are dimensionless coordinates

(9)
Along the body, for an inviscid flow, the slip condition is enforced,
--t --t

that is,

V.n =0

(10)

ule + viTI = 0

(11 )

This condition is enforced weakly by reducing the flux integral along any portion of
the body surface 8r b to a pressure integral:

The domain of study is discretized either by quadrangles to form a structured


mesh, or by triangles, possibly to form an unstructured mesh (see Figure 3). In all
cases, the nodes supporting the degrees of freedom are chosen to be the vertices, and
in particular, a row of such nodes is introduced along the body. An arbitrary node
is referred to by the index i, or equivalently, in the case of a structured mesh, by the
subscripts i,j. A dual finite-volume mesh is constructed by forming around each node
i, a control area, or finite-volume cell, Ai, by dividing by the medians the quadrangles

392 Computational Methods in Hypersonic Aerodynamics


into 4, or the triangles into 3, and connecting portions of these medians to close the
cell. In this way, a cell boundary is made of 4 interfaces, each one made of 2 segments
in the case of a structured quadrangulation, but the number of neighboring triangles
or interfaces is arbitrary in the case of an unstructured mesh (Figure 3 (b)).
Defining the cell average,

Wi

= ff.A;Wdxdy/Ai

(13)

a first-order explicit time-integration method for (6) is given by:

Ai

W!'+l - W!'

r.-

+ Lj hIB;j (F(wn)nx + G(Wn)ny) ds =

(14)

in which the summation on j is made over all the neighbors of i, or equivalently, over
all the interfaces AIBij between node i and a neighbor j.
To complete the definition of the numerical method, the spatial approximation is
defined by the procedure to evaluate the integral along, or the flux through an arbitrary interface. This is done to first-order and second-order accuracy in subsequent
subsections.
First-order numerical flux
A first-order spatial scheme is defined by the following approximation:

[_ . (F(W)nx
hIB;]

+ G(W)ny) ds ~ F(WI )7]x + G(WI )7]y

(15)

in which AIBij is a typical interface and

7}ij = (7]x) = [_ rt ds
7]y
hIB;]

(16)

is the integrated normal.


Upwinding is introduced to define the intermediate value WI in terms of the "left"
and "right" vectors Wi and Wj!iS. One way to realize this is to observe that for some
appropriate rotation matrix no,

F(WI)7]x

+ G(WI )7]y =

II7}ijlln o- 1 F(n oWJ)

(17)

and treat the problem one-dimensionally.


We have employed the van Leer flux-vector splitting

F(V)

= F+(V) + F-(V)

(18)

in which the precise definition of the vectors F+ (V) and F- (V) can be found in [22].
This leads to the following quadrature formula:
(19)
in which the positive part of the flux is evaluated upwind, and the negative downwind. Combining (14-19) yields the following first-order explicit method symbolically
represented by
W n+1 wn
i
n --t )
A i i ~(20)
+ "L...J." .T.(wn
'l'
i , Wj , 7] ij = 0
t
J

Computational Methods in Hypersonic Aerodynamics 393


in which cJ>, the so-called "numerical flux function", has the following form:

(21 )
This method is stable under a C.F.L.-like condition, and the (first-order accurate)
steady-state solution depends only on the spatial approximation. Before examining
how the spatial accuracy can be enhanced to second-order, an implicit time-integration
algorithm is constructed for stability and steady-state efficiency.
Implicit formulation
An implicit analog of (20) is

win+1 - wni + ""' cJ>(W!,+1 W!'+l r;*,,) =


tlt
L . J . ' ) , ./ .)

(22)

This equation is nonlinearly implicit, since the dependence of the numerical flux cJ> on
its first two arguments is nonlinear. Such algorithms are usually very robust. This
is why, for certain applications in which the equations to be solved are very stiff,
one such formulation is adopted at the cost of solving a set of nonlinear algebraic
equations at each timestep (which can be done for example, by Newton's iteration,
or any other fixed-point technique more specifically designed to the structure of the
associated quasi-linear system). Here, the applications considered, that is, Euler flows
of inert or dissociating air in the hypersonic regime, revealed only moderately stiff.
AS,a result, a linearized version of the above scheme is sufficiently robust (stable) in
practice. For this, one writes

iIi.(w!,+1
W!,+1
r;* .. ) =

,
)
"/')

'I'

cJ>(wn
wnj "r;*
)
i'
/ ij
+cJ>u(Wr, Wjn, Tlij).(Wt+ 1

wt)

(23)

+cJ>v(Wr, Wp, Tlij).(WP+1 - WP)


Knowing analytically the Jacobians,

(24)
allows us to express
cJ>u
{ cJ>v

= 111lllnii1 A+ (nou) no
= 111lllnii1 A- (no V) no

As a result, the (linearized) implicit discrete equation at node i is:

~~ ~cJ>v(Wr, Wp, 1lij) 8Wjn =


)

(25)

394 Computational Methods in Hypersonic Aerodynamics

~~ ~cI>(Wr, Wp, llij)

(26)

in which for any node k, the following notation has been employed:

"wnk -- W kn+1 _ wnk

(27)

Equation (26) is written in the so-called "delta form". It is a linear system of


equations in which the unknowns on the left-hand side are the increments 8Wr for all
k. The right member is the amount by which the vector Wr is incremented at step
n in the explicit method. The matrix contains the Jacobians of + and - parts of the
splitting employed to form the right member. Note that the steady-state solution is
entirely defined by this right member, the matrix in the left member therefore only
acts as a stabilizing preconditioner.
From a theoretical viewpoint, it can be shown for a linear hyperbolic equation,
W t + A Wx = 0, that the matrix obtained by an analogous construction is (similar to)
a diagonally-dominant matrix. Consequently, (26) can be solved by (Jacobi or GaussSeidel) relaxation (see e.g. [35]). In practice, best efficiency is achieved when only
solving such system to partial convergence by a few relaxation sweeps, and computing
a new timestep with updated Jacobians.
Large timesteps, not limited by the C.F.L. condition can usually be employed.
Very often, one successfully uses a different timestep at each gridpoint, calculated to
make the C.F.L. number approximately uniform over the grid, and increasing during the convergence process as the residual error decreases. In the limit of an infinite
timestep, the iteration identifies to Newton's method which has quadratic convergence
[36].
Quasi second-order extensions [20, 21]
Second-order steady-state solutions can be computed, by enhancing to second-order
accuracy only the approximation of the right member of (26), that is, the calculation of the numerical flux terms, not of their Jacobians. Such formulation combines
first-order (in the preconditioner) with second-order (in the explicit increment) approximation schemes. This slight inconsistency causes that iterative convergence can
at best be linear (and not quadratic) for an infinite timestep. In addition, a fullyupwind second-order scheme is pathological, and instead a half-fully upwind scheme
as in the construction by Fromm should be preferred [36].
Keeping in mind that such precautions should be made, the second-order extensions are based on the M.U.S.C.L. approach [19], in which one replaces the vectors Wi
and Wj in the role they play in the computation of the numerical flux, respectively by
vectors W ij and W ji , obtained from left and right linear extrapolations at the point
I, that is
cI>ij

= cI>(W;j, W}i, llij)

(28)

In practice, these extrapolations are made on the primitive variables p, u, v and p


that are "coded" afterwards in vectors of conservative variables. Hence, for example,
1-+

Qij

77

= qi + 2'VQi. Z)

(29)

Computational Methods in Hypersonic Aerodynamics 395


---t

in which q is any primitive variable, and V' qi is some approximation of the gradient
of q at node i [20, 21].
The following two techniques can be employed to perform the extrapolations.
(a) Extrapolation with slope limitation
In the PI-Lagrange approximation, gradients are constant by triangles. Thus, the
most natural way to extrapolate nodal values, is to first compute at each node i
---t

the average gradient of the variable V' qi' as the weighted average of the gradients
evaluated in the triangles surrounding the node, the weights being the areas of the
respective triangles; and secondly, to perform the extrapolation (29); thus,
area(Tj ).

(~) T
(j neighbor of i)

Lj area(Tj )

(30)

(yielding % and similarly, qji = qj - ~~j.0). In doing this, in the implicit formulation, while the preconditioner is a first-order upwind operator, the explicit update
(or right-hand side) is a "half-fully-upwind" second-order operator, since the corrections to the first-order upwind nodal values are based on centered gradients. This
combination is recommended in [36] for best iterative convergence.
In fact, the above second-order approximation is not robust enough for being
practical. In an attempt to construct a monotonic scheme, limitation is applied to
the gradient, or "slope", prior to the extrapolation, by the Min-Mod function: the
corrections

(~) T

.0 are calculated for all triangles Tj surrounding the node; if all


J

of these numbers are of the same sign, ~i'


is set equal to the one of smallest
modulus; otherwise it is set equal to zero, and the approximation is locally only
first-order accurate.
In conclusion, this construction results in a quasi-second-order half-fully upwind
scheme.
(b) Quasi- TVD scheme based on the upwind triangle
Another route to construct a half-fully upwind scheme, is to average the prediction
of a fully upwind scheme with that of a centered scheme. For f sufficiently small, the

point i -

f0 belongs to the same triangle, defined as the "upwind triangle" T jj .


---t

---t

fully upwind extrapolated value qij can then be computed by letting V' qj = V' qT' J in
(29). Hence, the corrections brought to the nodal value qi by the fully upwind scheme
and the centered scheme are, respectively:
1 ---t
-;--t
1
a = "2V'qT. J .lJ , b = "2(qj - qi)

(31)

The half-fully upwind scheme is then obtained by letting

= qi + Ave(a,b)

in which the symbol Ave stands for some averaging function.

(32)

396 Computational Methods in Hypersonic Aerodynamics


Again, if the averaging function is the arithmetic mean, the approximation lacks
robustness. Instead, in an attempt to construct a TVD-like scheme, the van-Albada
average [37] is employed:
if ab

>0

(33)

otherwise
where

is some small number to avoid zero divide.

Conclusion
The above construction provides an efficient implicit algorithm for the computation of
second-order accurate steady solution to the Euler equations. The following sections
are devoted to the extension of this algorithm to cases of dissociating air.
SIMULATION OF EQUILIBRIUM CHEMISTRY
Equilibrium model
An equilibrium model was originally constructed at AMD-BA by B. Stouffiet and M.
O. Le Ber [5, 24]. In this model, the gas is made of 5 species 0, N, NO, O2 and
N 2 , and 3 independent equilibrium equations are extracted from the non-equilibrium
dissociation model used by Park [9]:

20
2N
2NO

(34)

(These reactions are initiated by collision factors omitted in this writing.) The law of
mass action is thus expressed as follows:

Y?

_1

14

= mi
= m~

15

m5

I<2(T)
2

Y,2

1415

m4

Y,2
_2

I<1(T)

(35)

~I<3(T)
m4 m 5

in which the Y;'s for i = 1,2, .. ,5 are the mass fractions of 0, N, NO, O 2 and N2
respectively, and the mi's their respective molar masses; Y; = 8m;j8m where 8mi is
the mass of species i contained in a control volume 8V and 8m is the mass of the
mixture in this volume. In this way, the subscripts 1, 2 and 3 are associated with the
"chemical products", i.e. the species that are not present in the undisturbed flow, or
below a certain temperature To (To = 298 I<). For the equilibrium constants I<1 (T),
I<2(T) and I<3(T), we have employed the same expressions as in [9].
The mass fractions satisfy two algebraic equations. First, evidently:

(36)

Computational Methods in Hypersonic Aerodynamics 397


Second, since the local proportion of oxygen atoms relative to nitrogen atoms is (conserved thus) uniform:

Y2 Ya 2Y5
-+-+
-m5
m2
m3

(37)

79

Note that (35-37) can be solved for the mass fractions, once estimates for density and
temperature are known. Hence these equations can be thought as defining implicitly

Yi = f;(p, T).
Each individual species is treated as a perfect gas with constant specific heats
(CPi = 7/2R; for N 2 , O2 and NO and 5/2R; for Nand 0; R; = R/mi : gas constant).
The state equation has the form:
5 Yo
p = pRTE-'

(38)

;=1 mi

Conceptually at least, since Yi = !i(p, T), p = !(p, T). Thus, in some other methods,
equilibrium flows are computed by an ideal-gas solver solely modified by the change of
equation of state. (In this case, the functional relationship! is usually implemented
by tables, or "Mollier-diagram".)
Finally, one needs to relate energy, temperature and mass fractions. For this, the
specific enthalpy is first expressed:
5

h=

E
1; C
.=1

Pi

K8'!

+ .=1
~ 1; h? + E 1;
(~)'
.=3
exp T

(39)
-

where h? is the formation enthalpy of species i. The formation enthalpies of 0 and


N can be related to the dissociation temperatures of O2 and N 2 :

h~ = h~ = Rl 8~2 /2
{ hg = h~ = R28'k2/2

( 40)

8~ = 59500 K
{ 8'k: = 113 000 K

(41 )

in which

[25], and the formation heat of NO is


h~

= h~o =

2.99 106

J.kg- 1

( 42)

The last term in (39) corresponds to the vibrational energy of the molecules assumed
at equilibrium with the translational and rotational modes, and 8y is the characteristic
vibrational temperature found in [25] :

83 = 8 No
{

= 2 740 K

8~ =

8 02 = 2 270 K

8~ =

8 N2 = 3 390 K

( 43)

398 Computational Methods in Hypersonic Aerodynamics


As a result, the total energy per unit volume is

V2

E=p(h+T)-p

(44)

Note that this equation can be solved for the temperature, which appears nonlinearly,
once the mass fractions are estimated, all other properties being given by the solution
of the flow equations.
Explicit numerical scheme
The basic ingredient of the algorithm is the solver for the Euler equations, [26], described in the previous section. The code is modified to evaluate the pressure according to the equation of state for a mixture of perfect gases (38), and the parameter,
everywhere it appears as the ratio of enthalpy to internal energy,

,=-he

(45)

("equivalent", "efficient" or "pseudo"-,), where e is the internal energy per unit mass:

p E
V2
e=h--=--p
p
2

( 46)

Similarly, a pseudo-soundspeed

(47)
and a pseudo-Mach number

M=V/c

( 48)

are injected in the formulae of the flux-vector splitting in place of the usual properties
defined for ideal gas. Note that, is only needed when flux splitting is employed, since
it does not appear explicitly in the partial differential equations. Therefore, in any
case, the differencing scheme remains consistent with the complete set of governing
equations, and if the definition of the parameter, contains some arbitrariness, it however only affects the artificial viscosity of the numerical approximation. In addition,
for supersonic flow F+ = F, thus the artifice only affects the subsonic region where,
in our applications, it has been numerically verified that M ~ M.
These modifications being made, the flow equations are advanced forward in time
over one timestep by the explicit method. Then the temperature and the mass fractions are calculated by solving (35-38) iteratively (Newton's method). At convergence
the value of , is updated in preparation of the next timestep.
The process is continued until the steady state is achieved.
Implicit Formulation
From a conceptual viewpoint, adjoining an equilibrium model to the Euler equations
is equivalent to modifying the state equation p = f(p, T), since one can think of the
composition variables Y;'s as auxiliary variables that could be eliminated and replaced
by expressions of e.g. p and T (or p). In view of this, the implicit formulation given
by (26) is applicable; however, accounting for a modified state equation would, to be
rigorous, necessitate a different formal evaluation of the Jacobians A and B, and the

Computational Methods in Hypersonic Aerodynamics 399

question of defining adequate split forms A+, A -, B+ and B- should again be posed
(flux-vector splitting for equilibrium chemistry).
If such a process of differentiation with constraints is well founded from a theoretical stand, it is questionable that it yields a truly superior algorithm, compared
with the simplified implicit algorithm that one achieves by letting these Jacobians
have the usual formal expressions and evaluating them from the local equivalent '"Y
and sounds peed, which account for chemistry. Such a simplified implicit construction
is simple and has been revealed already to be very efficient.
Mesh adaption
The shock initiates the dissociation phenomena, and it may be useful that some
efforts be made to locally increase the spatial accuracy in this area. To capture such
a structure economically, if a structured mesh is employed, a one-dimensional mesh
adaption very similar to the "static rezone" of [27] can be applied.
In an explicit solution, every say 500 timesteps, a new mesh is constructed, the
solution interpolated onto it and the time integration proceeds. Technically, a sensor
of the local Mach gradient, or criterion function is computed along a given ray (11 =
constant). The following form revealed adequate:
w

where

= a + bm r + cm rr

- I~I18M I'
a;:-

(49)

- I~I18 MI
max

mr - max
r

mrr -

(50)

8r2

and a, band c are constants close to 1. Then the points are redistributed along the
ray so that

w=

w(r)dr

(51 )

rw

be linear in the computational radial coordinate (which is a node index), and precisely
not directly in r.
In fact, although the adapt ion is made one-dimensionally, a 2D smoothing of the
criterion w is performed to produce more regular meshes. This is done by solving two
scalar tridiagonal systems,

(52)
in which 08 and Or are second-difference operators in the circumferential and radial
directions and f = 5; then one uses the function Wi (r) instead of w( r) in the calculation
ofw.

This technique is very adequate to cluster cells in the strong shock. However,
if this technique is to be extended to Navier-Stokes computations, or applied to a
quasi "frozen flow" calculation where important phenomena occur in the vicinity of
the obstacle, it is not very effective to refine the mesh near the wall as necessary. This
can nevertheless be achieved by modifying the criterion function w( r) to include a
term proportional to IIcur! VII in (49) in order to detect sudden flow variations both
in the shock and the boundary layer.

400 Computational Methods in Hypersonic Aerodynamics


With unstructured triangulations, the mesh can be adapted to the shock and more
generally to one-dimensional structures by node movement or mesh enrichment by element division. See [28, Chapter IV] for a general presentation of such techniques,
and the following section on non-equilibrium flow for one example.
Numerical experiments
In these experiments, the basic explicit first-order method is employed to compute hypersonic flows about a blunt body made of a cylindrical portion followed by planes inclined of 15 degrees. The freest ream conditions are taken to approximately correspond
to the standard atmosphere at the altitude of 75 km (Poo = 2.52 Pa; Poo = 0.43 10- 4
kg/m3 ; Too = 205.3 K). The first two experiments are made with a (coarse) structured
mesh of 41 x 31 points. This mesh appears on the right side of Figure 7.
The cases of Figure 4 correspond to a freestream Mach number of 17.9. There,
the nonreactive perfect gas flow is compared with the chemically reacting gas flow.
The iso-Mach number contours, Figure 4 (a), demonstrate the important reduction
of the stand-off distance, and in fact, of the entire shock layer due to the dissociation
phenomena, that has the global effect of diminishing the parameter,. The temperatures, Figure 4 (b), are reduced by more than roughly 65 %. The mass fractions
plots, Figure 4 (c), indicate that the dissociation occurs immediately after the shock,
and then evolves rather smoothly with x. At this Mach number, O2 is completely
dissociated, NO appears only in minute amounts and thus 0 in in constant amount;
the reactional scheme reduces to an exchange between N2 and N. On the Mach number and temperature plots, Figure 4 (d), the reduction of the stand-off distance and
of the temperature distribution are more accurately identified. Note that with the
upwind solver, the flow is strictly constant upstream of the shock, and the solution
is oscillation-free. Note how little the effect of chemistry is on the Mach number
along the body. This observation deserves some attention. In the classical theory of
hypersonics for a non-reactive gas [29], the flow is invariant as the freest ream Mach
number increases ("Mach Number Independence Principle"). A close examination of
the Rankine-Hugoniot jump conditions through a normal shock reveals that the limit
Moo --+ 00 has the effect of the limit, --+ 1, which in this sense, is comparable to
the effect of chemical dissociation. For these reasons, the Mach number wall distributions presented in this article are altogether very close, for perfect or reactive gas,
independently of the freest ream Mach number.
On Figure 5, the effect of the freest ream Mach number is more precisely evaluated. The iso-Mach number contours, Figure 5 (a), are essentially invariant, but the
temperatures, Figure 5 (b), are somewhat more sensitive to Moo. For example, the
stagnation temperature is roughly augmented by 35 % when Moo increases from 15
to 25.
It has sometimes been alleged that the main features of a chemically dissociated
gas flow are essentially those of a frozen flow with, = 1.2, at least for certain engineering productions. We observed this fact on iso-Mach number contours, however,
the stagnation static temperature of such a frozen flow represented on Figure 6 by a
straight line departs noticeably from the data of Figures 4 and 5 reproduced here on
this plot. Hence, we do not confirm the allegement.
The mass fractions plots, Figure 5 (c), reveal that the strength of the dissociation

Computational Methods in Hypersonic Aerodynamics 401


varies with Moo but not its nature: a larger quantity of N2 is dissociated in N as Moo
increases. On Figure 5 (d), the three Mach number plots are essentially the same,
while the temperatures increase somewhat with Moo.
On Figure 7, the effectiveness of mesh adaption is demonstrated. Three calculations are made of the same flow (Moo = 17.9, equilibrium chemistry solver), with
three different meshes: (1) a coarse, essentially uniform mesh of 41 X 17 points, (2)
an adapted mesh of the same number of points, and (3) the fine but uniform mesh
of 41 x 31 points. The examination of the corresponding iso-Mach number contours
indicate that the sharpest shock is given by the adapted mesh calculation, and in fact
also at least cost.
In the remaining experiments, calculations of flows around a cylinder were made at
Moo = 20 with the finite-element extension of the code. In a first experiment, a coarse
mesh of 650 points is employed; the iso-Mach number contours are plotted on Figure
8 (a) to compare the first-order accurate approximation with the second-order. In
the second-order accurate solution, the shock is closer to the body since the artificial
viscosity has been reduced. In a second experiment, a fine mesh of 6140 points is
employed to compute a more accurate solution. Corresponding iso-Mach, iso-T, isoCp, iso-p, iso-Y N2 and iso-Y N contours are shown on Figure 8 (b-g) respectively. From
this, the stand-off distance is found to be approximatively 15 % of the radius.
Further descriptions of methodologies of this type can be found in [38, 39, 40]
along with the presentation of a variety of experiments related to equilibrium flow
computations.
SIMULATION OF NON-EQUILIBRIUM CHEMISTRY
As mentioned in the introduction, important non-equilibrium effects do occur in the
shock layer at high temperatures, when the characteristic geometrical length is of
comparable size with the chemical relaxation lengths. Hence, for a space shuttle,
non-equilibrium effects occur near the nose, and even more importantly near the
winglets where the radius of curvature is small (a few cm). There the temperature
needs to be controlled very cautiously, and this necessitates accurate calculations.
Non-equilibrium model
When T ~ 2500 K, diatomic species N2 and O2 react: dissociations, recombinations
and ionizations occur transforming the initial composition in a mixture of many compounds: N 2, O 2, NO, N, 0, NO+, 0+, N+, 0-, N-, e- (electrons), ... in proportions
that are functions of the temperature [2].
We are employing several models inspired from [30]. These models are relatively
simple and have been tested to give reasonable equilibrium compositions in the temperature range of concern [26]. Ionization is neglected. Five species are retained 0,
N, NO, O2 and N2 subscripted from 1 to 5 in this order. The most complete model

402 Computational Methods in Hypersonic Aerodynamics


includes 18 reactions:

O2 + X
N2 + X
NO + X
NO + 0
N2 + 0
N2 + O2

(Rt}

O+O+X
N+N+X
N+O+X
O2 + N
NO + N
NO + NO

(R 2 )
(R3)
(R4)
(Rs)
(Re)

(53)

X represents a collision factor which can be any species of the five present. Here, (R 1 ),
(R 2 ) and (R3) are dissociation reactions and (R4) and (Rs) are bimolecular exchange
reactions. The latter are the most important for the formation of nitric oxide NO in
air. Note that the reaction (Re) can be obtained by adding the reactions (R4) and
(Rs). However, contrasting with the equilibrium composition, in a nonequilibrium
phenomenon, the chemical equations do not have to be independent.
A simpler model can be obtained by only retaining the following 3 reactions:

20 + N2
NO + N
O2 + N

(54)

It should be emphasized that the species production from N2 and O2 is endothermic,


the reactions will thus result in a high temperature decrease.
Again, the composition of the mixture is defined by the values of the mass fractions
Y;.
The conservation of mass for species i is expressed by

ata (PY;) + div(pY; -V) = n

(55)

where ni is the mass production rate of species i (production term). The complete
expressions for these terms are explicated in Appendix I for the 3-reaction model; for
the 18-reaction model the expressions are similar but involve more terms. From a
mathematical standpoint, these terms include factors of the form:
constant

TS

exp ( -

~)

n:

(pY;)"i

revealing the very strong dependency on temperature and the nonlinear character of
the phenomenon. In addition, the discrepancy between the rate constants associated
with the different reactions has the effect of increasing the stiffness of the problem
(disparate timescales).
In our models we have considered a single temperature for the mixture. However, more complete models may includes besides the classical translation-rotation
temperature T, several vibrational temperatures for the diatomic molecules. Correspondingly, transport equations for the associated vibrational energies can be cast in
a mathematical form similar to (55), and require a similar numerical treatment. However, to be consistent, when vibrational relaxation is accounted for, the rate equations
yielding the precise expression for n, should also be modified. Hence, these models
could be rather more complicated to define f311.

Computational Methods in Hypersonic Aerodynamics 403


Ionization is also sometimes included in more complex chemical reactional schemes.
The first ion to appear is NO+ [25]. However, it amounts to less than 1 % as long
as T < 25000, which is the case if Moo < 25. Thus, for many applications in fluid
mechanics, ionization does not play an important role. However, the presence of ions
has significant consequences in the field of Hertzian communications.
To summarize, as in the equilibrium-chemistry calculations, we have used:
- global mass conservation, (36),
- global conservation of 0 and N atoms, (37),
- the equation of state for a mixture of perfect gases, (38),
- the expression of total energy in terms of flow variables (V, p, ... ), temperature,
T, and mass fractions, Y; (i = 1,,5) given by (39-44).
This allowed us to eliminate Y02 and YN2 and solve only three conservation equations:
i = 1,2,3 in (55).
Slip conditions are applied to (55) on wall boundaries; hence, as the divergence
term does not contribute to the body line integrals appearing in the global mass
conservation equation, such term does contribute either to the species conservation
equations. Initially, the inert gas composition (21 % O 2 , 79% N 2 ) is specified throughout the domain.
1-D non-equilibrium chemistry solver - scales and stiffness [32]
In order to get some insight on the physical scales associated with the chemistry
phenomena, and the related stiffness of the problem, one can simulate a situation in
which the aerodynamic flow can be simplified. Here one considers the one-dimensional
flow behind a strong normal shock.
If one assumes that the convection characteristic length of the flow is much larger
---+

than the characteristic relaxation length, the density P and the velocity V can be held
fixed to the values given by the jump conditions (Rankine-Hugoniot relations). In this
model, one can solve (55) for i = 1,2,3 for the mass fractions y;, the temperature
being related to the mass fractions via a constant total enthalpy equation applicable
to steady flow,

h+

V2

(56)

= constant

in which h is given by (44).


Thus the system to be solved is given by:

a
a
at (Pi) + ax (pN) =

Oi

(i = 1,2,3)

(V> 0)

(57)

in which the unknowns are the density functions Pi(X, t) = PY;(x, t), V is a constant
equal to the normal component of velocity behind a normal shock at a given freestream
Mach number Moo, and Oi is a nonlinear function of temperature T and Pi (i=1,2,3):

Oi = Oi(Pt, P2, P3, T)


The model accounts for 5 species and 18 reactions.

(58)

404 Computational Methods in Hypersonic Aerodynamics


These equations were solved by the linearized Euler implicit method based on a
2nd-order partially-upwind scheme explicitly, and a 1st-order upwind scheme implicitly:
(59)
in which

Pi)
( ~~ ,

(60)

and
(61)
is a 3 X 3 Jacobian matrix. In (59), \7 is the first-order backward-difference operator,
and 8~ is a linear combination of the central-difference operator and the second-order
backward-difference operator. They are defined by:
\7Wj

= Wj -

Wj-I

(62)
+ ~ (3Wj - 4Wj_1 + Wj-2)
In practice, the half-fully upwind scheme (/3 = t) is usually employed [28 Chapter II,
{

8~wj =

(Wj+! -

Wj-I)

32,36].
Remark 1: At the steady state, the P.D.E. reduces to the system

(VW)x

=n

(63)

Hence, the stationary solution can also be computed by space marching. Performing
this integration by Runge-Kutta 4 on a fine mesh yields the same solution without
noticeable difference in accuracy [32].
Remark 2: In a separate study [32],the eigenvalues of the Jacobian matrix n'(w) have
been computed numerically and found to have negative real parts. This is no surprise
since the integration along a streamline should lead to equilibrium conditions. Additionally, the diagonal dominance of the matrix preconditioner in (59) is augmented
by the chemistry term.
In the one-dimensional experiments, we have employed the nonequilibrium model
of [5]. The initial solution is uniform and corresponds to the known conditions behind
a normal shock. I
The results of the calculations for Moo = 15, 20 and 25 are shown on Figure 9
(a), (b) and (c) respectively, on which are given the plots vs x of normalized source
terms Fi = n;fpv (i = 1,2,3), mass fractions and temperature. As the Mach number
increases, from 15 to 20 and 25, the region near x = 0, where the source terms are
important, reduces noticeably from about 5 mm, to 0.5 mm and to 0.05 mm. In this
region, dissociation is abrupt and the temperature falls rapidly, giving an indication
on the stiffness of the problem. Beyond, the evolution to equilibrium composition is
1 The

vibrational temperature is accOlmted for to calculate the temperature.

Computational Methods in Hypersonic Aerodynamics 405

rather smooth, and occurs with a characteristic relaxation length much larger, say
about 5 or 10 cm.
In these calculations, the implicit method is unconditionally stable. (This was
tested for C.F.1. numbers up to 300,000.) The convergence history plot is given on
Figure 10 for C.F.L. numbers equal to 10, 1000, 100,000. For large timesteps, the
asymptotic convergence is equivalent to that of the sequence 2- n regardless of the
number of mesh intervals [28 Chapter II, 36].
Numerical algorithm for 2-D Euler equations
Extending the notations employed for ideal gas, the 2-D motion of an inviscid fluid
subject to non-equilibrium chemistry is governed by the following set of P.D.E.'s
written in weak-conservation law form:

Wt

+ oxF(W) + oy G(W)

= O(W)

(64)

in which now:

P
W=

pu
pv
E
PI

P2
P3

F(W) =

pu
pu 2 +p
puv
u(E + p)

G(W) =

pv
puv
pv 2 + p
v(E + p)

o
o
o

Pt U

Pt V

O(W) =

0
!1 t

P2 U
P3 U

P2 V
P3 V

!13

O2
(65)

We refer to the previous sections for the expression of the source terms Oi (i
Again:

= 1, 2, 3).

- pressure is given by the equation of state for a mixture of perfect gases, (38),
- total energy (per unit volume), E, is given in terms of the flow variables, the
temperature and the mass fractions by (39-44), and
- the global conservation of 0 and N atoms allows us to express the mass fractions
of O2 and N2 in terms of the others.
It is instructive to examine the gross effect of the geometricallengthscale Rb on the
physical nature of the typical blunt-body flow. For this, denote the specified solution
vector in the undisturbed flow by Woo. The steady-state solution is then defined by
the following set of equations:
Rb-l (~eF(W) + ~f1G(W)) = O(W)
W=Woo
{ ufe+ v ff1=O
+ thermodynamic relations

over D
along rin
along rb

(66)

408

Computational Methods in Hypersonic Aerodynamics

In this way, the splitting applied to the species convection terms accounts for all three
eigenvalues associated with the linearization of the Euler equations and reinforces the
coupling between the fluid-motion equations and the chemistry equations.
Also, if in some region of the flow n -+ 0, the solution Y =const. can be found
numerically exactly. This differencing was found to be much superior to the "donorcell" approximation.
Remark 3: In the finite-element context, to derive (73) one applies "mass-lumping"
on the diagonal terms to maintain the O.D.E. structure in the discrete equation.
In this algorithm, a 3 X 3 system is solved at every node, the rest being explicit.
This inversion is facilitated by the diagonal dominance of the matrix [Remark 2 above].
However, scales greatly disparate in the entries necessitate careful programming of the
inversion routine.
In this form, the method is first-order accurate in both time and space, but well
suited to the extension to second-order spatial accuracy by the M. U.S.C.L. procedure.
We now present the results of a few experiments made with this first-order method
and the 41 X 31 finite-volume mesh of Figure 7.
In a first experiment, the freestream Mach number and the nose radius of curvature
are set equal to: Moo = 17.9, Rb = 1 m. Three chemical reactional schemes are
compared accounting for 3, 6 and 18 reactions respectively [31]. On Figure 11 (a)
and (b), the iso-Mach number and iso-temperature contours are essentially identical.
However, the temperatures are found to be noticeably lower in the calculation of
the 18-reaction model, which is believed to be more accurate. Examining the mass
fractions plots, Figure 11 (c), we see that the more complete model predicts a greater
dissociation of O2 and a larger production of atomic nitrogen N. Comparing these
results with those obtained in the equilibrium-chemistry case, Figure 4, the novelty is
apparent in the temperature plots that exhibit a peak just after the shock triggering
the dissociation.
On Figure 12, we examine the influence of the nose radius of curvature Rb on the
solution of the 18-reaction model, at Moo = 17.9. The computed cases are defined by
Rb = 103 m, 1 m and 10-3 m, a range of 6 orders of magnitude. The first evidence is
that the non-equilibrium algorithm has been able to sharply compute three smooth solutions indicative of near-equilibrium, typical non-equilibrium and near frozen regime
respectively without particular adjustment. As Rb increases, the size of the shock layer
diminishes and approaches the extent of the equilibrium solution. As a result, the level
of the temperatures decreases significantly; also, the pattern of the iso-temperature
contours is modified. It can be seen on the mass-fractions plots, Figure 12 (c), that
in the near-frozen flow N2 is almost not dissociated at all, and O2 only a little. In
the intermediate case, products of dissociation appear in noticeable amounts. The
near-equilibrium case is almost identical to the equilibrium case with respect to mass
fractions (compare with Figure 4 (c)); with respect to the temperatures (Figure 12
(d)), the discrepancy is only visible by the presence of a small peak immediately after
the shock. This can be explained as follows: from the jump conditions satisfied by the
exact solution of the hyperbolic system modeling this non-equilibrium flow, we know
that the mass fractions should be continuous through the shock ("frozen shock") and
consequently, along the axis of symmetry, where the shock is normal, the tempera-

Computational Methods in Hypersonic Aerodynamics 409


ture jump is the same as in the inert-gas solution; hence, the exact non-equilibrium
solution is expected to contain a very high peak of temperature immediately behind
the shock; in the numerical solution, a truncation of this peak is visible.
The first case, Rb = 103 m, is found very near to the equilibrium case of Figure 4.
Hence, we verify that in the limit of large radii, the non-equilibrium flow solver has
the capability to produce a similar answer to the equilibrium solver, as expected from
a previous argument.
The previous experiment is repeated for a freestream Mach number of Moo = 25
with the same geometries. It appears on Figure 13 that the same qualitative phenomena occur. However, the dissociation is more important at the higher Mach number,
and, for example, the frozen-flow conditions should be obtained with smaller geometries.
"Weakly-Coupled" Implicit Formulation
In the previous section, an explicit time integration of the Euler equations coupled
with a finite-rate chemistry model has been constructed based on a fractional-step
approach in which, at each timestep, the Euler equations are first advanced assuming
the chemical composition is frozen but non-uniform ("Euler sub-step"), and then the
species convection equations are advanced, after updating the temperature from the
current value of energy ("Chemistry sub-step").
The same fractional-step approach can be applied to construct an implicit algorithm. For this, in the "Euler sub-step", one employs the usual implicit algorithm,
but the Jacobians (as in the implicit solver for equilibrium flow) assume values locally
dependent on the equivalent 'Y and soundspeed. In the "Chemistry sub-step", the
source term is implicited and linearized, and since the differencing of the convection
terms is exactly modeled on the differencing of the continuity equation, scalar (i.e.
proportional to the identity matrix) Jacobians appear. This gives:
W~+l -

A,

tl.t

w!'

'" +
(w~+l
+!' -;:;+
.. )
'
"w
')
"')
n

"L
j

neighbor of

n
= A," (n~ + nm(w
+! -

w n ))

(77)
One could solve this nonlinear implicit equation by some iterative process, e.g. Newton's method. Instead we prefer to employ the following approximate linearization of
the flux term, which is inspired from (75)):

(78)

where own+! = wn+l - w n . These terms are substituted in (77) and grouped by nodes
(summation on j); if N3 is the number of nodes, this implicit formulation results in a
linear system for the N3 unknown 3-component vectors ow;'+! (i = 1, ... , Ns), where
the right-hand side is the usual explicit update. The system is typically solved by
some 5 to 10 Gauss-Seidel sweeps.

410 Computational Methods in Hypersonic Aerodynamics


In the literature, such fractional-step algorithm is sometimes referred to as a
"weakly-coupled" scheme. In a more straightforward formulation (not necessarily
more efficient in terms of cost), usually referred to as a "strongly-coupled" approach,
one would form pointwise, for 2-D flow, the complete 7x7 Jacobian of the 7 P.D.E.'s.
Here instead, at each node, one has formed first a 4x4 Jacobian of the Eulerian part;
this Jacobian was simplified, since in taking the partial derivatives, chemistry has been
assumed frozen; then, one has formed a 3x3 Jacobian of the chemistry equations, assuming the velocity (non-uniform) but frozen. Thus, in the implicit preconditioner
of the "delta-form", some cross-term derivatives have been neglected. However, one
should stress that this simplification by no means implies a simplification of the mathematical model solved at steady-state. It only possibly affects the robustness of the
implicit procedure, but the cost per timestep is much lower because the simplified
Jacobians are much less costly to evaluate.
The implicit solver can operate with much larger timesteps (unlimited in the simpler cases), usually controlled by the norm of the residual (or right-hand side), and the
iteration number n. Typically, the CFL number is gradually increased from 0.4 (usual
limit of the explicit method) to 20 or 50, and this results in a reduction of the number
of the necessary iterations by 40, and of the computation time by 15. A thorough
investigation of the iterative properties of this implicit scheme can be found in [41, 42].
Second-Order Schemes
Second-order schemes [42, 43] are constructed in a manner similar to the ideal-gas case
(see section on Basic Euler Solver). As previously, one employs extrapolation with
slope limitation, or constructs a quasi- TVD scheme based on the upwind triangle
technique. The critical issue, however, is the choice of variables that is made to
carry the extrapolation to the interface, before "coding" into conservative variables.
For inert gas, one employs the primitive variables p, u, v and p. For chemicallyreacting gas, the necessary control of the temperature leads to replace p by T which
is efficient but costly, or by the internal energy (here per unit mass) t, as suggested
by Montagne et al. [44], which is more economical. This alternative is adopted here.
The extrapolation of the chemistry variables is performed on the mass fractions.
Before giving an example of a second-order implicit calculation of a non-equilibrium
flow, a discussion is presented to describe the pattern of the stagnation point region
in an inviscid non-equilibrium external flow.
The Stagnation-Point Issue
In this section, a theoretical approximate solution to the species-convection equations
applicable in the neighborhood of a stagnation point is obtained from [45] and permits
to discuss the difficult issue of validation of a non-equilibrium flow solver.
For this, consider the steady, non-equilibrium flow over a blunt-body, and denote
by 8 the curvilinear abscissa along the stagnation streamline (with s < 0 upstream
the stagnation point, and s > 0 along the body) (Figure 14 (a)).
At the stagnation point (8 = 0) the (steady) mass conservation equation degenerates into the incompressibility condition
~

div( V) = 0

(79)

Computational Methods in Hypersonic Aerodynamics 405


rather smooth, and occurs with a characteristic relaxation length much larger, say
about 5 or 10 cm.
In these calculations, the implicit method is unconditionally stable. (This was
tested for C.F.L. numbers up to 300,000.) The convergence history plot is given on
Figure 10 for C.F.L. numbers equal to 10, 1000, 100,000. For large timesteps, the
asymptotic convergence is equivalent to that of the sequence 2- n regardless of the
number of mesh intervals [28 Chapter II, 36].
Numerical algorithm for 2-D Euler equations
Extending the notations employed for ideal gas, the 2-D motion of an inviscid fluid
subject to non-equilibrium chemistry is governed by the following set of P.D.E.'s
written in weak-conservation law form:

Wt

+ ax F(W) + ay G(W) = n(W)

(64)

in which now:

W=

pu

pu
pv
E

p
pu
puv
u(E p)

F(W) =

G(W) =

pv
puv
pv 2 p
v(E p)

PIU
P2 U
P3 u

PI

P2
P3

+
+

o
o
o

n(W) =

P1 V
P2 V
P3 V

01

O2
03
(65)

We refer to the previous sections for the expression of the source terms Oi (i
Again:

= 1,2,3).

- pressure is given by the equation of state for a mixture of perfect gases, (38),
- total energy (per unit volume), E, is given in terms of the flow variables, the
temperature and the mass fractions by (39-44), and
- the global conservation of 0 and N atoms allows us to express the mass fractions
of O2 and N2 in terms of the others.
It is instructive to examine the gross effect of the geometricallengthscale Rb on the
physical nature of the typical blunt-body flow. For this, denote the specified solution
vector in the undisturbed flow by Woo. The steady-state solution is then defined by
the following set of equations:
Rb-I (~eF(W)

+ ~7IG(W)) = n(W)

W=Woo
{ ufe+ v f 7l =0
+ thermodynamic relations

over D
along fin
along fb

(66)

408 Computational Methods in Hypersonic Aerodynamics

In this way, the splitting applied to the species convection terms accounts for all three
eigenvalues associated with the linearization of the Euler equations and reinforces the
coupling between the fluid-motion equations and the chemistry equations.
Also, if in some region of the flow n ..... 0, the solution Y =const. can be found
numerically exactly. This differencing was found to be much superior to the "donorcell" approximation.
Remark 3: In the finite-element context, to derive (73) one applies "mass-lumping"
on the diagonal terms to maintain the O.D.E. structure in the discrete equation.
In this algorithm, a 3 X 3 system is solved at every node, the rest being explicit.
This inversion is facilitated by the diagonal dominance of the matrix [Remark 2 above].
However, scales greatly disparate in the entries necessitate careful programming of the
inversion routine.
In this form, the method is first-order accurate in both time and space, but well
suited to the extension to second-order spatial accuracy by the M.U.S.C.L. procedure.
We now present the results of a few experiments made with this first-order method
and the 41 X 31 finite-volume mesh of Figure 7.
In a first experiment, the freestream Mach number and the nose radius of curvature
are set equal to: Moo = 17.9, Rb = 1 m. Three chemical reactional schemes are
compared accounting for 3, 6 and 18 reactions respectively [31]. On Figure 11 (a)
and (b), the iso-Mach number and iso-temperature contours are essentially identical.
However, the temperatures are found to be noticeably lower in the calculation of
the 18-reaction model, which is believed to be more accurate. Examining the mass
fractions plots, Figure 11 (c), we see that the more complete model predicts a greater
dissociation of O 2 and a larger production of atomic nitrogen N. Comparing these
results with those obtained in the equilibrium-chemistry case, Figure 4, the novelty is
apparent in the temperature plots that exhibit a peak just after the shock triggering
the dissociation.
On Figure 12, we examine the influence of the nose radius of curvature Rb on the
solution of the 18-reaction model, at Moo = 17.9. The computed cases are defined by
Rb = 103 m, 1 m and 10- 3 m, a range of 6 orders of magnitude. The first evidence is
that the non-equilibrium algorithm has been able to sharply compute three smooth solutions indicative of near-equilibrium, typical non-equilibrium and near frozen regime
respectively without particular adjustment. As Rb increases, the size of the shock layer
diminishes and approaches the extent of the equilibrium solution. As a result, the level
of the temperatures decreases significantly; also, the pattern of the iso-temperature
contours is modified. It can be seen on the mass-fractions plots, Figure 12 (c), that
in the near-frozen flow N2 is almost not dissociated at all, and O2 only a little. In
the intermediate case, products of dissociation appear in noticeable amounts. The
near-equilibrium case is almost identical to the equilibrium case with respect to mass
fractions (compare with Figure 4 (c)); with respect to the temperatures (Figure 12
(d)), the discrepancy is only visible by the presence of a small peak immediately after
the shock. This can be explained as follows: from the jump conditions satisfied by the
exact solution of the hyperbolic system modeling this non-equilibrium flow, we know
that the mass fractions should be continuous through the shock ("frozen shock") and
consequently, along the axis of symmetry, where the shock is normal, the tempera-

._-

- .. -

----- - -

-----------

Computational Methods in Hypersonic Aerodynamics 409

ture jump is the same as in the inert-gas solution; hence, the exact non-equilibrium
solution is expected to contain a very high peak of temperature immediately behind
the shock; in the numerical solution, a truncation of this peak is visible.
The first case, Rb = 103 m, is found very near to the equilibrium case of Figure 4.
Hence, we verify that in the limit of large radii, the non-equilibrium flow solver has
the capability to produce a similar answer to the equilibrium solver, as expected from
a previous argument.
The previous experiment is repeated for a freest ream Mach number of Moo = 25
with the same geometries. It appears on Figure 13 that the same qualitative phenomena occur. However, the dissociation is more important at the higher Mach number,
and, for example, the frozen-flow conditions should be obtained with smaller geometries.
"Weakly-Coupled" Implicit Formulation
In the previous section, an explicit time integration of the Euler equations coupled
with a finite-rate chemistry model has been constructed based on a fractional-step
approach in which, at each timestep, the Euler equations are first advanced assuming
the chemical composition is frozen but non-uniform ("Euler sub-step"), and then the
species convection equations are advanced, after updating the temperature from the
current value of energy ("Chemistry sub-step").
The same fractional-step approach can be applied to construct an implicit algorithm. For this, in the "Euler sub-step", one employs the usual implicit algorithm,
but the Jacobians (as in the implicit solver for equilibrium flow) assume values locally
dependent on the equivalent 'Y and soundspeed. In the "Chemistry sub-step", the
source term is implicited and linearized, and since the differencing of the convection
terms is exactly modeled on the differencing of the continuity equation, scalar (i.e.
proportional to the identity matrix) Jacobians appear. This gives:
w~+1 - w!'

ilt

'"'

A..

L...J

(w~+l w~+l
-;:;+ .. )
3
'" '3

'"

= A' (n~
I , + n'~(w~+l _
''I

w n,

))

neighbor of i

(77)
One could solve this nonlinear implicit equation by some iterative process, e.g. Newton's method. Instead we prefer to employ the following approximate linearization of
the flux term, which is inspired from (75)):

lilt:
[T')

(w.n

-;:;+ .. )]

".")

1 '"

n+l

uw

[11l~

T.)

(wn
)

-;:;+ .. )]
."J

(78)
1 '"

n+l

uw
)

where 8w n +l = wn+l - w n. These terms are substituted in (77) and grouped by nodes
(summation on j); if Ns is the number of nodes, this implicit formulation results in a
linear system for the Ns unknown 3-component vectors 8wi+l (i = 1, ... ,Ns ), where
the right-hand side is the usual explicit update. The system is typically solved by
some 5 to 10 Gauss-Seidel sweeps.

410 Computational Methods in Hypersonic Aerodynamics


In the literature, such fractional-step algorithm is sometimes referred to as a
"weakly-coupled" scheme. In a more straightforward formulation (not necessarily
more efficient in terms of cost), usually referred to as a "strongly-coupled" approach,
one would form pointwise, for 2-D flow, the complete 7x7 Jacobian of the 7 P.D.E.'s.
Here instead, at each node, one has formed first a 4x4 Jacobian of the Eulerian part;
this Jacobian was simplified, since in taking the partial derivatives, chemistry has been
assumed frozen; then, one has formed a 3x3 Jacobian of the chemistry equations, assuming the velocity (non-uniform) but frozen. Thus, in the implicit preconditioner
of the "delta-form", some cross-term derivatives have been neglected. However, one
should stress that this simplification by no means implies a simplification of the mathematical model solved at steady-state. It only possibly affects the robustness of the
implicit procedure, but the cost per timestep is much lower because the simplified
Jacobians are much less costly to evaluate.
The implicit solver can operate with much larger timesteps (unlimited in the simpler cases), usually controlled by the norm of the residual (or right-hand side), and the
iteration number n. Typically, the CFL number is gradually increased from 0.4 (usual
limit of the explicit method) to 20 or 50, and this results in a reduction of the number
of the necessary iterations by 40, and of the computation time by 15. A thorough
investigation of the iterative properties of this implicit scheme can be found in [41, 42].
Second-Order Schemes
Second-order schemes [42, 43] are constructed in a manner similar to the ideal-gas case
(see section on Basic Euler Solver). As previously, one employs extrapolation with
slope limitation, or constructs a quasi- TVD scheme based on the upwind triangle
technique. The critical issue, however, is the choice of variables that is made to
carry the extrapolation to the interface, before "coding" into conservative variables.
For inert gas, one employs the primitive variables p, u, v and p. For chemicallyreacting gas, the necessary control of the temperature leads to replace p by T which
is efficient but costly, or by the internal energy (here per unit mass) t, as suggested
by Montagne et al. [44], which is more economical. This alternative is adopted here.
The extrapolation of the chemistry variables is performed on the mass fractions.
Before giving an example of a second-order implicit calculation of a non-equilibrium
flow, a discussion is presented to describe the pattern of the stagnation point region
in an inviscid non-equilibrium external flow.
The Stagnation-Point Issue
In this section, a theoretical approximate solution to the species-convection equations
applicable in the neighborhood of a stagnation point is obtained from [45] and permits
to discuss the difficult issue of validation of a non-equilibrium flow solver.
For this, consider the steady, non-equilibrium flow over a blunt-body, and denote
by 8 the curvilinear abscissa along the stagnation streamline (with 8 < 0 upstream
the stagnation point, and 8 > 0 along the body) (Figure 14 (a)).
At the stagnation point (8 = 0) the (steady) mass conservation equation degenerates into the incompressibility condition
--+

div( V) = 0

(79)

Computational Methods in Hypersonic Aerodynamics 411


Consequently, in the neighborhood of this point, the incompressible-flow approximation is accurate. In this approximation, the density is constant along the streamline,
and the velocity profile is linear:

p = p(s) = a constant,
{ V = V(s) = k lsi

(SO)

where V(s) is the velocity modulus (Figure 14 (b)).


The (steady) species convection equations can be written in the following nonconservative form:
--t _
p V .VY = O(p, T, Y)
(Sl )

Yaf

in which Y = (Yt,}'2,
and 0 = (OIl O2 , 03)T, and the continuity equation has
been used. If one projects these equations on the stagnation streamline, and employs
the incompressible-flow approximation, one gets:
pk

dY(s)

lsi d;- = O(T(s),Y(s))

(S2)

in which the formal dependence of 0 on p has been suppressed since the density is a
constant.
This equation implies that at the stagnation point (s = 0), either (a) dY(s)jds is
infinite, or (b) 0 = 0, or (a) and (b) hold simultaneously. We will admit that dY (s) j ds
may become infinite as s -+ 0- , that is, when the limit is taken normal to the wall, but
that it remains finite as s -+ 0+, that is, when the limit is taken tangentially to the
body surface. As a consequence of this assumption, the condition in (b) is satisfied,
and it states that equilibrium conditions are realized at the stagnation point:

00 = 0

(S3)

where the subscript 0 refers to s = O. In fact, note that the traveling time of a fluid
particle along the streamline from some abscissa So < 0 (in the shock layer) to Sl = 0
(stagnation point) is

7(So)

1
so

ds

-V(s) =

(S4)

00

if the incompressible-flow assumption is made. Thus, from a physical point of view, it


is not surprising that the particle has achieved equilibrium conditions when reaching
the stagnation point.
By virtue of the conservation of total specific enthalpy,
k2
2S2

+ Cp(Y)T + hoY + HVIB(Y, T) =

H oo ,

(S5)

the temperature can be thought as an implicit function of S2 and Y:

(S6)
Note that this implies in particular that

(aT)
as

= (a~

as

2S)

= o.
0

(S7)

412 Computational Methods in Hypersonic Aerodynamics


An approximate solution to the O.D.E. (82) can be obtained after linearization of
the source term about stagnation values (equilibrium conditions):

n - no = (an)
ay

(Y _ Yo)
0

+ (an) (T - To) + ...

aT

(88)

The second term of this equation can itself be expressed using the expansion:

T - To = (aT)
as (s _ 0) + (aT)
ay (Y - Yo) + ...
0

But we have already seen that


terms yields:

(89)

no = 0 and (aT I as)o = 0, thus, neglecting second-order


(90)

where n~ is a 3 x 3 matrix that can be called the "total Jacobian" of the source term
with respect to the mass fractions, since it also reflects the variation of temperature
due to a variation of composition:

, (an an aT)
no = ay + aT ay o'

(91)

As a result, the set of linearized O.D.E.'s writes:

n'
pkl:1 (Y -

dY

ds

Yo)

(92)

Introducing the diagonalization of the matrix n~,


n~

=T

A T- 1

(93)

one assumes (and this has been verified numerically, [32, 45]) that the eigenvalues of
the diagonal matrix A are real and negative. This means that the equilibrium state
is attractive. Consequently, the eigenvalues of the matrix
A
pk

D=--

(94)

are positive, and their definitions indicate that they are Damkholer numbers since 1I k
is a convective timescale and the elements of Alp the inverses of chemical relaxation
timescales.
Two subcases are examined in the solution of the linearized equation (92).
For s < 0, ahead of the stagnation point in the shock layer, the solution is easily
found to be:
Y(s) = Yo + T (_s)D C
(s < 0)
(95)
where C is a vector of constants, and for any real number a > 0, the following notation
is defined:

(96)

Computational Methods in Hypersonic Aerodynamics 413


Hence, the derivative along the streamline is

dY
ds

= _T

(_S)D-I C

(s < 0)

(97)

in which I denotes the identity matrix.


Past the stagnation point (s > 0) and along the body surface, the form of the
general solution is
(s > 0)
(98)
Y(s) = Yo + T (st D C'
where C' is a vector of constants. The only bounded solution is obtained for C' = 0,
which yields:
(s > 0).
Y(s) = Yo
(99)
For a typical non-equilibrium hypersonic flow of air, the numerical computation
of the Damkholer numbers gives [45]:

Dn < 1
{ D22> 1
D33> 1

(100)

where the eigenvector associated with Dn is closely related to the mass fraction of
N, that is, to the dissociation of N 2 Consequently, upstream of the stagnation point
(s < 0), a continuous solution is found that has infinite left gradient dY/ds(O-) = 00.
Downstream of the point (s > 0), along the body, the solution (of the linearized
equation) is constant. This solution is sought to indicate the local behavior of the
solution of the nonlinear equation, (82).
The singularity can be shown [45] to disappear if the freest ream Mach number or
the size of the obstacle is large enough. However, for practical cases, this is not the
case, and thus in the neighborhood of the stagnation point, the mass fraction of N in
particular, and the temperature usually have singular behaviors of the type indicated
of Figure 14 (c) and (d).
As a result, the wall distributions of temperature and mass fraction of N given
by a non-equilibrium inviscid flow solver are truncated according to the mesh size
near the wall. This explains why equilibrium conditions are usually not found at the
stagnation point by a conventional discretization method, unless extremely fine mesh
size is locally employed. In particular, the temperature may, in some severe cases, be
over-estimated by several thousand Kelvin, as observed during a recent workshop on
hypersonic flows [40]. In fact, a sequence of solutions obtained over finer and finer
meshes would converge extremely slowly at the wall.
In the experiments of Figures (12) and (13), equilibrium conditions have been
approached in the case of a large body (Rb = 103 m). The interest of those experiments was lying in the fact that they demonstrated that the discrete solution had
the correct qualitative behavior as the Damkholer number increased. However, in the
light of the present theoretical analysis, it appears that such experiments could not
be quantitatively correct with respect to stagnation values, or in the rough estimation
of the value of Rb for which stagnation values are close to equilibrium.
Although the importance of the singularity is limited by the fact that it is inherent to the "unphysical" inviscid-flow model, the numerical analyst should be aware of

414 Computational Methods in Hypersonic Aerodynamics


it for its impact on validation of numerical methodologies, and also if an Euler-type
algorithm is designed to be coupled with a reactive boundary-layer calculation. For
more details on these aspects, see [45].
An example: The Flow over a Double-Ellipse
To illustrate the calculation over an unstructured grid of a hypersonic reactive flow
by the implicit second-order accurate method, we reproduce here some of the results
presented at an International Workshop on Hypersonic Flows [46].
The test case considered here is the external, inviscid non-equilibrium flow over a
double-ellipse geometry. The semi-axes of the large ellipse are 60 cm and 15 cm, and
of the small ellipse, 35 cm and 25 cm. The freestream Mach number is 25 and the
angle of attack 300
The employed mesh (Figure 15) contains some 14,000 points and has been obtained from a structured initial mesh by two enrichments by element division in shock
regions, identified by a criterion estimate based on the local Mach number gradient.
Figure 16 provides the iso-Mach number contours on which the detached strong
shock and the canopy shock are clearly visible. The solution by the quasi-second-order
scheme is smooth but sharply resolved.
Finally, Figure 17 provides the wall distribution of temperature, T, and mass
fraction of atomic nitrogen, YN . In the standard result, the value found at the stagnation point is near 9000 K, that is about 3000 K superior to the equilibrium value
at the same pressure (5750 K), whereas YN is under-estimated. This illustrates the
discussion of the previous section, in which it was concluded that in a conventional
computation, the temperature profile would normally be truncated near the stagnation point. A post-processed result is also shown from [45]; it is obtained by assuming the pressure distribution correctly determined by the standard procedure and by
computing using this assumption, the stagnation properties as equilibrium values; a
space-marching technique is then applied to the steady-state species-convection equations and projected momentum equation to obtain the wall distributions. As a result,
the wall distribution of temperature is basically shifted downward by 3000 K. Again,
this post-processed solution is not believed to be close to some physical solution that
could be computed by the solution of an appropriate Navier-Stokes problem. It only
illustrates the chemical boundary layer in the theoretical inviscid non-equilibrium flow
and permits the appreciation of the difficulty inherent to this model.
CONCLUDING REMARKS
In conclusion, we gather on Figure 18 the Mach number plots and the temperature
plots obtained for a freest ream Mach number Moo = 25, and in the following three
cases:
(1) ideal gas flow,
(2) equilibrium-chemistry flow,
(3) non-equilibrium flow, Rb = 1 m, I8-reaction model.
The following stand-off distances can be measured from the Mach number plots:

Computational Methods in Hypersonic Aerodynamics 415

(1) D/Rb
(2) D/ Rb
(3) D / Rb

~
~
~

0041,
0.14,
0.25.

These distances are numerically over-estimated. For example, in the ideal gas case,
some theory of hypersonic flow [29, Fig. 6-6] predicts that D / Rb ~ 0.39. The exact
reasons for this discrepancy remain an open question; the usage of a first-order, hence
dissipative, scheme over an insufficiently fine mesh has certainly played a role.
Nevertheless, one can observe that chemistry has the effect of reducing the standoff distance of approximately 50 % in this case. The temperature levels are also
reduced by about 70 %, but the patterns are noticeably different in the equilibrium and
non-equilibrium cases. In the latter case, the temperature peak immediately behind
the shock falls abruptly to a stagnation temperature superior to the equilibrium value;
the temperature along the body decreases smoothly.
The shape of the body and the freestream conditions being fixed, the temperature
plot in the non-equilibrium flow case diminishes monotonically from the inert gas
solution (Rb -+ 0) to the equilibrium-gas solution (Rb -+ 00) as Rb increases.
The calculation of a steady non-equilibrium flow over a double-ellipse by a secondorder implicit method over a mesh adapted by several successive enrichments has
demonstrated the efficiency and robustness of the upwind approximation coupled
with adaptive unstructured grids.
Finally, these methods can be extended to flows governed by more complex models. For example, the extension to flows in thermal non-equilibrium can be found
in [47]; a natural extension of the approximation to viscous flows governed by the
Navier-Stokes equations is given in [43J, while various possible models of the viscous
terms are provided and assessed in [48J.
ACKNOWLEDGEMENTS
The author would like to express his gratitude to his collaborators and particularly to
those who co-authored many related publications with him, and whose contributions
to the field of hypersonics have been essential in preparing this chapter. Namely R.
Abgrall, N. Botta, M. C. Ciccoli, A. Dervieux, 1. Fezoui, N. Glinsky, E. Hettena, A.
Merlo, C. Olivier, M. Passalacqua and M. V. Salvetti at INRIA Sophia Antipolis, and
M. Mallet, B. Stouffiet and J. Periaux at Dassault Aviation.
The author would also like to acknowledge the cooperation with R. Brun, M.
Imbert, D. Zeitoun and their colleagues at Departement Milieux Hors Equilibre, Universite de Provence, for constant stimulating interactions.
REFERENCES

1. J.-A. DESIDERI, N. GLINSKY, E. HETTENA, Hypersonic Reactive Flow Computations, Computers and Fluids Vol. 18, No.2, pp. 151-182, 1990.
2. J. D. ANDERSON, Jr., Modern Compressible Flow with Historical Perspective,
McGraw-Hill Book Company (1982).

416 Computational Methods in Hypersonic Aerodynamics


3. C. P. 11, Hypersonic Nonequilibrium Flow past a Sphere at Low Reynolds Numbers, AIAA Paper No. 74-173, AIAA 12th Aerospace Sciences Meeting, Washington, D.C./Jan.30-Feb.1, 1974.
4. R. W. MACCORMACK, Numerical Solution of the Interaction of a Shock Wave
with a Laminar Boundary Layer, Lectures notes in Physics, Springer-Verlag, New
York, 8 151-163 (1971).
5. J. V. RAKICH, H. E. BAILEY, C. PARK, Computation of Nonequilibrium
Three-Dimensional Inviscid Flow over Blunt-Nosed Bodies Flying at Supersonic
Speeds, AIAA Paper No. 75-835, 1975.
6. A. W. RIZZI, H. E. BAILEY, Split Space-Marching Finite-Volume Method for
Chemically Reacting Supersonic Flow, AIAA Journal 14, 621-628, May 1976.
7. J. D. RAMSHAW, 1. D. CLOUTMAN, Numerical Method for Partial Equilibrium Flow, J. comput. Phys. 39,405-417 (1981).
8. D. K. PRABHU, J. C. TANNEHILL, Numerical Solution of Space Shuttle Orbiter Flowfield Including Real-Gas Effects, J. Spacecraft 23 264-272 (1984).
9. C. PARK, On the Convergence of Chemically Reacting Flows, AIAA Paper 850247, AIAA 23rd Aerospace Sciences Meeting, Jan. 14-17, 1985/Reno, Nevada.
10. P. COLELLA, H. M. GLAZ, Efficient Solution Algorithms for the Riemann Problem for Real Gases, J. comput. Phys. 59, 264-289 (1985).
11. P. GLAISTER, An Approximate Linearized Riemann Solver for the Euler Equations For Real Gases, J. comput. Phys. 74,382-408 (1988).
12. S. EBERHARDT, G. PALMER, A Two-Dimensional, TVD Numerical Scheme
for Inviscid, High Mach Number Flows in Chemical Equilibrium, AIAA Paper
86-1284, AIAA/ASME 4th Joint Thermophysics and Heat Transfer Conference,
June 2-4, 1986/ Boston, Massachussetts.
13. J. P. DRUMMOND, M. Y. HUSSAINI, T. A. ZANG, Spectral Methods for Modeling Supersonic Chemically Reacting Flowfields, AIAA Journal, 24 1461-1467
(1986).
14. J. L. MONTAGNE, Utilisation d'un Schema Decentre pour la Simulation d'Ecoulements Non Visqueux de Gaz Reels it l'Equilibre, La Recherche Aerospatiale,
433-441 1986-6.
15. J. L. MONTAGNE, H. C. YEE, M. VINOKUR, Comparative Study of HighResolution Shock-Capturing Schemes for a Real Gas, NASA Technical Memorandum 100004, July 1987.
16. J. L. MONTAGNE, H. C. YEE, G. H. KLOPFER, M. VINOKUR, Hypersonic
Blunt Body Computations Including Real Gas Effects, NASA Technical Memorandum 10074, March 1988.

-------

Computational Methods in Hypersonic Aerodynamics 417


17. G. V. CANDLER, R. W. MACCORMACK, The Computation of Hypersonic
Flows in Chemical and Thermal Nonequilibrium, Paper No. 107, Third National
Aero-Space Plane Technology Symposium, June 2-4, 1987.
18. N. BOTTA, M. PANDOLFI, M. GERMANO, Nonequilibrium reacting hypersonic flow about blunt bodies: numerical prediction, AIAA Paper 88-0514, AIAA
26th Aerospace Sciences Meeting, January 11-14, 1988/ Reno, Nevada.
19. B. van LEER, Towards the Ultimate Conservative Difference Scheme III.: Upstream-Centered Finite Difference Schemes for Ideal Compressible Flow, J. comput. Phys. 23, 263-275 (1977).
20. A. DERVIEUX, Steady Euler Simulations using unstructured meshes, von Karman Institute for Fluids Dynamics, Lectures Series 1985-04, (1985), published in
Partial Differential Equations of Hyperbolic Type and Applications, G. Geymonat
(Ed.), World Scientific, (1987).
21. F. FEZOUI, Resolution des Equations d'Euler par un Schema de van Leer en
Elements Finis, INRIA Report No. 358, January 1985.
22. B. van LEER, Towards the Ultimate Conservative Difference Scheme I. The
Quest of Monotonicity, Lectures notes in Physics, Springer Verlag, New-York,
18 163-168 (1973).
23. 1. FEZOUI, H. STEVE, Decomposition de Flux de van Leer en Elements Finis,
INRIA Report No. 830, April 1988.
24. B. STOUFFLET, M. O. LE BER, private communication.
25. W. G. VINCENTI and C. H. KRUGER, Jr., Introduction to Physical Gas Dynamics, R. E. Krieger Publishing Co., Malabar/ Florida, Inc., (1982).
26. J-A. DESIDERI, E. HETTENA, Numerical Simulation of Hypersonic Equilibrium-Air Reactive Flow, INRIA Report No. 716, august 1987.
27. B. L. LARROUTUROU, Adaptive Numerical Methods for Unsteady Flame Propagation, Combustion and Chemical Reactors, G. S. S. Ludford Ed., Lectures in
Applied Math., 24 (2), 415-435, AMS, Providence (1986).
28. J-A. DESIDERI, A. DERVIEUX, Compressible Flow Solvers Using Unstructured
Grids, von Karman Institute Lecture Series 1988-05, March 7-11 (1988).
29. W. A. HAYES, R. F. PROBSTEIN, Hypersonic Flow Theory, second edition,
Volume I: Inviscid Flows, Applied Mathematics and Mechanics Volume 5A, Academic Press, New York and London, 1966.
30. Y. P. RAIZER, Y. B. ZEL'DOVICH, Physics of Shock Waves and High-Temperature Hydrodynamics Phenomena, Academic Press, New York, 1965.
31. R. BRUN, P. COLAS, P. GUBERNATIS, D. ZEITOUN, Practical PhysicoChemical Models for High-Speed Air Flow-Field Computations, Rapport Final
Hermes R.D.M.F. 86, June 1989.

418 Computational Methods in Hypersonic Aerodynamics


32. J-A. DESIDERI, L. FEZOUI, N. GLINSKY, Numerical Computation of the
Chemical Dissociation and Relaxation Phenomena behind a Detached Strong
Shock, INRIA Report No. 774, December 1987.
33. J-A. DESIDERI, L. FEZOUI, N. GLINSKY, E. HETTENA, J. PERIAUX, and
B. STOUFFLET, Hypersonic Reactive Flow Computations around Space-ShuttleLike Geometries by 3-D Upwind Finite Elements, AIAA Paper 89-0657, AIAA
27th Aerospace Sciences Meeting, Jan. 9-12, 1989/Reno, Nevada.
34. B. LARROUTUROU, L. FEZOUI, On the Equation of Multionent Perfect or
Real Gas Inviscid Flow, Nonlinear Hyperbolic Problems, Carasso, Charrier, Hanouzet and Joly Eds., Lecture Notes in Mathematics 1402, Springer- Verlag, Heidelberg, 1989.
35. H. STEVE, Schemas Numeriques Linearises Decentres pour la Resolution des
Equations d'Euler en Plusieurs Dimensions, These, UniversiU de Provence AixMarseille I, 1988.
36. J.-A. DESIDERI, P. HEMKER, Analysis of the Convergence of Iterative Implicit
and Defect-Correction Algorithms for Hyperbolic Problems, INRIA Report No.
1200, 1990.
37. B. VAN LEER, Computational Methods for Ideal Compressible Flow, Von Karman Institute for Fluid Dynamics, Lecture Series 1983-04, Computational Fluid
Dynamics, March 7-11, 1983.
38. E. HETTENA, Schemas Numeriques pour la Resolution des Equations des Ecoulements Hypersoniques it l'Equilibre Chimique, These, Universite de Nice-Sophia
Antipolis, 1989.
39. N. BOTTA, M.-C. CICCOLI, J.-A. DESIDERI, L. FEZOUI, N. GLINSKY, E.
HETTENA, C. OLIVIER, Reactive Flow Computations by Upwind Finite Elements, Proceedings of the Workshop on Hypersonic Flows for Reentry Problems,
Part I, Antibes, France, January 22-25, 1990, Springer- Verlag, J.-A. Desideri,
R. Glowinski, J. Periaux Eds., 1991, to appear.
40. J.-A. DESIDERI, Some Comments on the Numerical Computations of Reacting
Flows over the Double-Ellipse (Double-Ellipsoid), Proceedings of the Workshop
on Hypersonic Flows for Reentry Problems, Part I, Antibes, France, January 2225, 1990, Springer- Verlag, J.-A. Desideri, R. Glowinski, J. Periaux Eds., 1991,
to appear.
41. M.C. CICCOLI, L. FEZOUI, J.-A. DESIDERI, Methodes Numeriques Efficaces
pour les Ecoulements Hypersoniques Non Visqueux Hors Equilibre Chimique, La
Recherche Aerospatiale, to appear.
42. N. GLINSKY, L. FEZOUI, M.C. CICCOLI, J.-A. DESIDERI, Non-Equilibrium
Hypersonic Flow Computations by Implicit Second-Order Upwind Finite Elements, Proc. of the Eighth GAMM-Conference on Numerical Methods in Fluid
Mechanics, Notes on Numerical Fluid Mechanics, Vol. 29, Vieweg, Braunschweig
1990.

Computational Methods in Hypersonic Aerodynamics 419


43. N. GLINSKY, Simulation Numerique d'Ecoulements Hypersoniques Reactifs
Hors-Equilibre Chimique, These, Universite de Nice-Sophia Antipolis, 1990.
44. J. 1. MONTAGNE, H. C. YEE, M. VINOKUR, Comparative Study of HighResolution Shock-Capturing Schemes for Real Gases, NASA TM 10004, 1987.
45. M. V. SALVETTI, J.-A. DESIDERI, Inviscid Non-Equilibrium Flows in the
Vicinity of a Stagnation Point, INRIA Report, to appear.
46. M. V. SALVETTI, M. C. CICCOLI, J.-A. DESIDERI, Non-equilibrium Inviscid
and Viscous Flows over the Double-Ellipse by Adaptive Upwind Finite-Elements,
Workshop on Hypersonic Flows for Reentry Problems, Part II, April 15-19, 1991
- Antibes, France (INRIA Publication).
47. A. MERLO, R. ABGRALL, J.-A. DESIDERI, Calcul d'Ecoulements Hypersoniques de Fluide Non Visqueux en Desequilibre Chimique et Vibratoire, Proc.
Journees d'Etudes sur les Ecoulements Hypersoniques, Roscoff, France, 22-24
Octobre, 1990.
48. M. V. SALVETTI, M. PASSALACQUA, Influence of the Physical Modelling
of Viscous Terms on Hypersonic Flow Computations, INRIA Report 1493, to
appear.

420 Computational Methods in Hypersonic Aerodynamics

APPENDIX I: Production rates for the 3-reaction model


The present calculation follows the derivation in [31].
The 3-reaction model is written as follows:

20 + N2
(101 )
NO + N
O2 + N
These reactions are cast into the following symbolic form in which appear the stoichiometric coefficients:
'"' 0''-' A
'"' 0'''-' A
(102)
L.J .)
L.J .)
In these notations, the subscript i refers to a species, and the subscript j = 1,2,3 to
the reaction (R j ).
The forward and reverse reaction rate constants associated with the reaction (R j )
(j = 1,2,3) are denoted kfj(T) and krj(T) respectively, and this allows us to define
the following functions of temperature and composition:
Jj = kf(T)
J

p' )171;j
)1711;j
II ( -'
- kr(T) II (P'
-'
m.
m.
J

-i=i,

i=ir

(103)

in which i l is the index of any species appearing on the left of the reaction (R j ) and

ir on the right.
The production rates are then expressed as follows:
3

O -- m '"'
L.J (0'"".)

0"")
. )J.
)

(104)

j=l

In particular:

01 =

m1

(2J1

O2 =

m2

(J2

03
04

J 1 = k/l (T)

J2

J3 )

+ J3 )

= m3 (J2 - J3 )
= m4 (-J1 + J3 )

05 =
where:

-m5

(105)

(J2 )

(~:) (~5J - kq (T)

(:::J 2(~5 )

J 2 = kJ2(T)

(~5J (~lJ - kr2(T) (~3J (~2J

= kfa(T)

(~3J (~lJ - kr3(T) (~:) (~2J

J3

(106)

Consequently, the definitions are completed by the formulae for the reaction rate
constants in terms of temperature:
(107)

Computational Methods in Hypersonic Aerodynamics 421

where the constants Gj ,


by the equation:

Sj

and!? can be found in [9]. Defining equilibrium constants

(108)

in which Z = 10000/T, while the constants Aij can also be found in [9], the reverse
rate constants are then given by:
(109)

422 Computational Methods in Hypersonic Aerodynamics

Moo 1

Figure 1: Flow pattern ahead of a hypersonic vehicle

out

out

Figure 2: Computational domain for blunt- body flow problem

Computational Methods in Hypersonic Aerodynamics 423

->

TJ

ij

J
A

Figure 3: Nodes, cells and interfaces


(a) discretization by quadrangles
(structured mesh)

(b) discretization by triangles


(possibly unstructured mesh)

424 Computational Methods in Hypersonic Aerodynamics


Figure 4: Comparison between Euler flows at Moo = 17.9
(2) chemically reacting gas
(equilibrium)

(1) perfect gas

(a) iso-Mach number contours

(b) iso-temperature contours

Tmax

= 13975 K

Tmax

= 4866 K

Computational Methods in Hypersonic Aerodynamics 425

Figure 4 (end): Comparison between Euler flows at Moo = 17.9

(c) mass fractions (equilibrium flow)


o.70~--~--~--~~

~~

-r ~:~:.~:~J~

o.ss
~oso

OO.4S

Go.40
~o.3S

'" 0.30

~~~~i"","""

~:~

0.30
C.2:I

),.

O.OS

0.llJL.0-.4-.0-.2----'0-"'C"O..,-2-..o-.4-..o-.6:--..o-.I~+1.0

O.ooL--~~-------"

0.4

-0.2

..0.2 ..0.4

..0.6 ..0.1

+1.0

(d) Mach number and temperature

(2) chemically reacting gas


(equilibrium)

(1) perfect gas

T(K)

l'iOll

11

11

10

10

lam

rr

j
T

.~l

Ii!
-0.4

-01

+01

+04

+0.6

+0.8

+1.0

-0.'

-01

+01

1'iOll

+0' +0.6

(x=O corresponds to the stagnation point)

+0.8

+1.0

426 Computational Methods in Hypersonic Aerodynamics


Figure 5: Effect of the freest ream Mach number (equilibrium chemistry)

Moo = 15

Moo

= 20

Moo

= 25

(a) iso-Mach number contours

(b) iso-temperature contours

Tmax

= 4276 K

Tmax = 5161 K

Tmax

= 5785 K

Computational Methods in Hypersonic Aerodynamics 427

Figure 5 (end): Effect of the freestream Mach number (equilibrium chemistry)

Moo

= 15

Moo

= 20

Moo = 25

( c) mass fractions
OIOC=:::;:T:::=::r:==~q
0.75~
0.70

z 0.65

0.10
0.7 l~
0.70
zO.65

<0.55

~ 0.60
(j Ol l

" 0.50

" 0.50

E060

I..

<

~o.4 l

:0.45

~O.40
1 0.15
O.JO
0.25

Io.J l

OJ0
0.25
O.lO
0.70
0.6l
060
O.ll

O.JO
0.25
O.lO
0.70
0.6l
0.60
OJl

O.70,------,-----.:==="'rl
0.65
060
0
0.55

1:~6
---- N

I~:.:. :.~~]6

QO.'l

gO.'l

t040

to.4O
0.30

., 0.30

~OJl

O.SO

III

ii

.'

,
'.

~OJl

~O.Jl

.,o.JO

I:~::::::.~6

0.50
CO.l
040

O.SO

V-

~O.40

~O.4 0
::1 OJ l

0.2DL-----'--------'

010
0.7 It------0.70
z 0.6l
060
< O.l l
!f 0lO
'" 0.4 l

~0.25_ .................. .

~0.25_ .......................................

~ O.2St----~.............. __ ...................... .

Io.lO
O.ll
0.10
O.OS

10.lO
O.ll
0.10
O.OS
0.00

10.lO
O.ll
0.10
O.OS
0.00

0.00L....--i.....c:.'""-_~====:;

.0.,

0.1

Hl.l

Hl.'

Hl.6

Hl.I

.l.0

------ll.4

-ll.2

to.2

.0.4 to.6 to.8 tl.O

-ll.'

~2

to.2

to.,

to.6

Hl.I

+1.0

(d) Mach number and temperature

Macib

T(K)

i,

Macth

f(K)

l5'

llOOl

3).

:n

1=

rI'I

s:

lOOl

_I

~ ..

[
.o. .ol

j
1=

"It

+0.2

+0...

+0.6

.08 .ID

llOOl

jllOOl

lJl

1=

"
10

IOf

l5

l5!

T(I<)

MacI

1
<l.'

.o.2

t
0

.0.2

to.4

-I
+0.6

lOOl

lOOl

+OJ

.ID

...1
.o. .ol

(x=O corresponds to the stagnation point)

+01

..04

.0.6

+0..8

+1.0

428 Computational Methods in Hypersonic Aerodynamics

T(K)

30000

- - - - - perfect gas, 'Y = 1.4


perfect gas, 'Y = 1.2

25000

o equilibrium gas

20000

15000
10000
o

5000

................
..

'

125

625

Figure 6: Stagnation-point temperature in terms of freest ream Mach number squared

Computational Methods in Hypersonic Aerodynamics 429

Mesh

Iso-Mach contours

Figure 7: Effect of Mesh adaptation (equilibrium chemistry)

430 Computational Methods in Hypersonic Aerodynamics

Figure 8: Finite-element calculations over a cylinder

Moo = 20
Coarse-grid calculations (650 points):

(a.l) iso-Mach contours; first order

MIN

= 0.25 ; MAX = 20 ; Ll = 0.25

(a.2) iso-Mach contours; second order

MIN

= 0.25 ; MAX = 20 ; Ll = 0.25

Fine-grid calculations (6140 points):

(c ) iso-Temperature contours

(b) iso-Mach contours

MIN

= 0.25 ; MAX = 20 ; Ll = 0.25

MIN

= 400 K;

MAX

= 5000 K;

Ll

= 200 K

Computational Methods in Hypersonic Aerodynamics 431

Figure 8: (end) Finite-element calculations over a cylinder

Moo = 20

Fine-grid calculation (cont.):


(d) iso-Cp contours

MIN

= 0 ; MAX = 1.9 ; .6. = 0.1

(e) iso- p con tours

MIN

(f) iso-YN2 contours

MIN

= 0 ; MAX = 0.19 ; .6. = 0.01

= 1 ; MAX = 15.5 ; .6. = 0.5


(g) iso-YN contours

MIN

= 0.58 ; MAX = 0.76 ; .6. = 0.01

432 Computational Methods in Hypersonic Aerodynamics

Figure 9: I-D non-equilibrium chemistry solver

(a) Moo = 15
F
50.

,
\
\

40. \
,
\

\.

30'

\\F1

20. r

\. "'"

'.'.
".

1O.

r/:~-- --~------->::"~c,'>______>
.

0.0

---0.2
0.4

0.6

0.8

.- 1.0 '" x(em)

SOURCE TERMS - (Yi)", = Fi(Y, T)


I::.

l.r

Y;

0.9 [
0.8

0.7f~------------------------------------------

0.6 _

0.5 ~
0.4,
0.3 0.2 \ 02
___________ 0
1
0 '.0 l~. . -.-..-.... . . . .. _-..-_-.--.-_..-.-.--__- - - - - - - . - J..NO
~._=~~~~~"
N:> x( em)
0
O.
2.
4.
6.
8.
10.
MASS FRACTIONS
30000J T(K)

~\~
. . . .~. ~
:.: : -"

_____ . __

25000
20000
r
15000r

100001~
5000

~======================l..,Tc>

L.

O.

2.

T ma",

4.
6.
8.
TEMPERATURE
= 10324. K T min = 6137. K

10.

x( em)

Computational Methods in Hypersonic Aerodynamics 433


Figure 9 (cont.): 1-D non-equilibrium chemistry solver
(a) Moo = 20
F

600j

48f'
360.~

F1
240. " ,

., .....

120..

-----------

O.oF:("-~

0.2

0.1

0.6

0.8

>

x(mm)

1.0

SOURCE TERMS - (Yi)., = Fi(Y, T)

l.oj

Y;

0.9,
0.8 eN

o.t

0.6r -------------------------------___ _
0.5.

OAt

0.3, 0

~:~l:;~~:=-=====---------==--===--~

0.0

r:----=:- .

O.

2.

30000L T(K)

No" x( em)

4.

6.

8.

10.

MASS FRACTIONS

25000
20000
15000
10000

- - - - - - - - - - - - - -__________ T

5000'--__~____~____~____~__~'__<> x( em)
O.
2.
4.
6.
8.
10.
TEMPERATURE
Trna., = 18185. K Tmin = 7349. K

434 Computational Methods in Hypersonic Aerodynamics


Figure 9 (end): I-D non-equilibrium chemistry solver

(a) Moo = 25
F
8000
\F2
6400

4800'

3200
....
....
........

0>

x(mm)

0.1

SOURCE TERMS - (li) .. = Fi(Y,T)

l.o~
0.9_

Y;

0.8l

0.7-t'N

0.6

o,sl

o4 t
---..----..--- N
. I'->~.=-~.::-.-- -------------------------------.
0.3,

0.2~9~---0.1,
0.0r-O.
A

~:!I

------..-0

NO
2.

4.

6.

8.

10.

x(cm)

MASS FRACTIONS

T(K)

15000

10000~_ _ _ _ _ _ _ _ _ _ _ _
50ooL,
,~x(cm)
T

0,

2.

Trna ..

4.

6.

8.

TEMPERATURE
T min = 7919. K

= 28291. K

10.

Computational Methods in Hypersonic Aerodynamics 435

iterations. n.
50

25

75

-2

\ \ / C.F.1. number = 10

-v

\
\

-5

-;;;

103 and 10 5

\
\

1
\

6;

.....
0

\
\

'--'

-8

\
\

\
\

\
\

CJ)

\
\

kl

'

-I::

<Il

...
....;j

-11

\
\

\
\

2- n

\
\
\
\

-14

\
\

\
\
\

\
\

I.
"I..

.II .....

:.

"1"'-v\.n..~V\""'I..:,~
\

Figure 10: Convergence history of the implicit I-D chemistry solver

436 Computational Methods in Hypersonic Aerodynamics

Figure 11: Comparison between nonequilibrium chemical models


Moo = 17.9 and Rb = 1 m
3 reactions

6 reactions

18 reactions

(a) iso-Mach number contours

(b) iso-temperature contours

T max = 8278. K

Tmax

= 8185. K

Tmax

= 7715. K

Computational Methods in Hypersonic Aerodynamics 437

Figure 11 (end): Comparison between nonequilibrium chemical models


1\-Joo = 17.9 and Rb = 1 m
3 reactions

6 reactions

18 reactions

(c) mass fractions

~~t----\

n~r---'---'-:==~"r1

L,---

0.101
z 0.61
060
COlI

!
~

OAS
<0.40
l0.31
0.30

:Il

E060
<0.5 I

:l!=O""~

;<0.50

:a.4Sf
~ OJOi

: 0.4 I
~ 0.40
~O.J I

o.:!l
0.~,L-

0.20'-------'---------'

_ _- L_ _ _ _ _ _

0'70r----r---r===n~

070,---,---.-:===n'"r1
061
0.60
0.11

0.30
0.25
0.20
0.70,r----,----,===nn

06l1
060

~.~

Ojsl

0.11

~ o.sol

~ 0.50

I:::::~:]6

~ o~sl
;:nJOI

20.41
(jO.40
<OJI
f!'
- 0.30

~OJSI

; o.sol

'"~ 0.25

lO.20
0.11
0.10
0.05
0.00

0611
0.601

lOJ!1
0...10'

0.25

:.:l.

0.7
0.70
zO.6I

070 1

~03!1

~050

0.80

l~ V-

Q.~

<Q~I

lO'~1
11
0. 1
nlo
0.4

0.2

j... .

,.'

0.11

nlo

.._..._.._.. : .. : ..

n05
.; <'~":
:.:~::
0.00,L---'""---:-.....;;.;.:.;.;.=-""";,;,;j
+C.2

+C..

+C.6

+C.I

-43.4

.1.0

0.05

0.00 -43.4

-43.2

;<\:~: . .----::_.:_.~.::_.;.~.:.:_a ..

+C.2

+C.4

+C.6

+C.I

.1.0

(d) Mach number and temperature

1lOh

T(K)

-:

II

"f

]1-

1-

:laxD

'----_--..[

t~~_

10\

'U~'-'~-"""~'
-0.4

.01

+0.2

+0.4

..0.6

+0 I

\.

+ ID

.(14

.0.2

--

+0.2

+0.4

+0.6

..oj . . 1.0

.0.4

(x=O corresponds to the stagnation point)

-01

+<>1

+04

+0.6

...oj

+1.0

438 Computational Methods in Hypersonic Aerodynamics


Figure 12: Effect of nose radius on nonequilibrium flows at Moo = 17.9
18 reactions model

(a) iso-Mach number contours

(b) iso-temperature contours

Tmax

= 5584. K

Tmax

= 7715.

Tmax

= 10979. K

Computational Methods in Hypersonic Aerodynamics 439

Figure 12 (end): Effect of nose radius on nonequilibrium flows at Moo = 17.9


18 reactions model

(c) mass fractions

OIOf==:::;-T---c:=:=::::xq
0.7l

nlOr----r---r--~~

0.10
0.7l
0.70
zO.6l
C
C O.bO
O..ll
<
<!O 0.50
:0.45
~ 0.40
:t OJl
0.30
O.ll

0.7l
070
zO.6l

0.70
z0.6l
C ObO
COll
<
<!OO.5O
~O.4j
~O.40
~0.35
0.30
O.ll

C ObO

E
< O.ll

i!' OJ()
; 0.45
< 0..0

l: O.Jl
0.30

0.25

O'lOL..--....L--------'

O.20L---....L--------'

0.20

070r---.---~=====n~

n70r----;---~=====n~

0.70
o.6l
0.60
O..ll

~.~

I~:~.:.~.~J6

0.6l
O.bO

O.ll

I~::::::::~0
..............

nll

~0.50
~O.4l

Z0.50

~o.so

~0.4l

:: 04l

~O.Jl

<O.3l

~ 030

i!'

H2O
O.Il

-0.30
~ O.2S
l:0.lO
nIl
0.10

'"

'f----------

~:

< o.JS

.. 0.30
:r. 0.25
~ O.lOr---, ... " .............................. .

'<0.25_ ......

..

E~40

t 0.40

t040

~:.~ ~.~.::~
I:gb

r----..

O.OS

.......................................
0.00L..{)~.4---,.{)~1""-0:--~+ll-:'.1.c.+ll-:...;.4"';{)'''':'-6"';{).':",:='-'.1.0

0.00'----&0-'--------'
.{).4 .{).l
0 ;{).l +ll.4 +ll.6 +ll.! LO

(d) Mach number and temperature


'j'T(K)

11l<Dl

",TOO

nK)

1
1

1l<Dl

I
illn))

II

II

10~

I I

Ii

.... :

;.laxlO

ri~

j
10

~!l<Dl
i

~.

-04

.02

+0.2

to.4

..0.6

+0.8

.1D

.0."

(x=O corresponds to the stagnation point)

-dl

.oJ +0.4

+0.6

-'

.0.1

+1.0

440 Computational Methods in Hypersonic Aerodynamics

Figure 13: Effect of nose radius on nonequilibrium flows at Moo


18 reactions model

= 25

(a) iso-Mach number contours

(b) iso-temperature contours

T max

= 6690.

Tmax

= 11505. K

Tmax

= 20092.

Computational Methods in Hypersonic Aerodynamics 441

Figure 13 (end): Effect of nose radius on nonequilibrium flows at Moo = 25


18 reactions model
10- 3 m

(c) mass fractions

O.7l

nW~==~-r----IC====~L

0.75

0.70

0.70

z 0.6l

0. W'

r:==;-:---C:::=3!:1
z n6l

z 0.6l
0.60

CO.ll

g060

CO.ll

~Oll

;< OlO
~O.4j

11: 0.50

;<0.50

<0.40
~OJS

0.45
II: 0.40

~OJl

;0.45
~o.40

0.30

0.30

OJO

0.25
0.20'------'-----------'

0.20'---------'---------------'

~ 060

<

<

III

l l ....
,..

0.25
0.20'---------'--------------'
0.70,-------,-------r:===-"

:~::::::~~6

0.6l

~i.~

IOJl

;~r----~i-----~I~~~:=~:~.~~~.~~~~~b'

[j 0.40
~OJl
':'n3O
'"

O.ll

~o.so

OJO

lmc-' I . . . . . . . . . . . . . .
".:

~ 025

O.ll
0.10
0.1ll

....

/I

!l:nJO
I

~O.25r---i

~:::.::::: ~u

0.4l

<OJ5

I 0.20

~~

t 0.40

'.

O70r-----..------,r-===_7Vl"lU
..

o.ll
,,0.4l

Q.l5

' ,

0.40
OJl

OJO
0.25

~~~---------I
..../
....................................

~~

nooIL.7"0.4'--.O~1-"-0:--:<d""2-~-:.""4-~-:-.6:--<d~i::--'+LO

0.00 -0.4

-0.2

~.2 ~.4

+0.6

~.i

LO

noo -0.4

~~~.2- ~.; ,o~:O';-:LO

-01

(d) Mach number and temperature

Mach:..

f(K)

:nl
t

IIxm

j
1

1=

"I ~I
I~
"I !~J

'j
.0.4

4.2

lL
0

-Kl.2

~..

-~=

+0.6

+OJ

+ID

-0.4

(x=O corresponds to the stagnation point)

-0.2

+0.2

+04

+0..6 +CJ +1.0

442 Computational Methods in Hypersonic Aerodynamics

p( s) = constant

shock wave

\
I
streamline
s

(a) Configuration

(b) Incompressible-Flow Model

(c) Local behavior of a mass fraction

(d) Local behavior of temperature

Figure 14: Inviscid Non-Equilibrium Flow Analysis in a


stagnation-point neighborhood

Computational Methods in Hypersonic Aerodynamics 443


Figure 15: Twice refined mesh for double-ellipse computation (14388 points)

444 Computational Methods in Hypersonic Aerodynamics

MIN = 0.25

MAX = 25.00

DLTA = 0.25

Figure 16: Iso-Mach number contours around double-ellipse

Computational Methods in Hypersonic Aerodynamics 445

10

STANDARD

~~~/~~.O
6

POST -PROCESSED

T
4

o~--~----~----~----~--~----~----~----~--~

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0.1

0.2

POST-PROCESSED

STANDARD

~:::

L... ..... ~

2~--~--~----~--~----~--~--~~--~--~

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0.1

0.2

x (m)
Figure 17: Wall distributions of temperature T (
(YN X 10)

1000 K) and atomic nitrogen

446 Computational Methods in Hypersonic Aerodynamics

(a) Mach number plots

(b) Temperature plots

--------

2!1~ ---------------;-.-.-.-.-.,

15000

:" ......
10000

10

~..I

........

~'"
-0.45

-0.4

-0.35

-0.)

-0.2.5

-0.2

-0.15

-0.1

-0.05

-0.1

- - - perfect gas flow


-- - - -- - - non-equilibrium flow
_. _. -. - equilibrium flow

(x=O corresponds to the stagnation point)

Figure 18: Comparison between Euler flows at Moo

= 25

-O.M

Chapter 12:
Computational Aerothermodynamics for 2D and
3D Space Vehicles
J. Hauser, J. Muylaert, H. Wong, W. Berry
European Space Research and Technology Center,
P. O. Box 299, 2200 A G Noordwijk, The Netherlands
ABSTRACT
This paper starts with an overview of current hypersonic activities in Europe: the Hermes spaceplane, the Titan probe Huygens, the Mars lander and studies on airbreathing
propelled future launchers. The special physical effects of high speed flows are first
discussed qualitatively and then scaling laws as well as relevant parameters, characterizing perfect gas, equilibrium and nonequilibrium flow are derived. After that, the
mathematical formulation of the governing physical equations is given, including the
physical submodels. In the next section the important role of supercomputers and of
the new massively parallel systems is discussed, presenting a picture of the future role
of this new architecture on aerospace computing. The numerical solution technique
is presented for multi-species Navier-Stokes equations, emphasizing the role of flux
linearization, to retain positivity. The section on grid generation gives a brief description of the package Grid* which is a toolbox for generating multiblock surface and
volume grids for arbitrary bodies. Grid adaptation and interactive grid generation
are also discussed. The last section depicts simulation results for 2D and 3D bodies,
starting with a N-S solution for a 2D hypersonic ramp followed by computations for
an axissymmetric flared cone. The 3D cases include a multiblock double ellipsoid,
and finally calculations for the Hermes Space Plane at Mach 21 are presented. The
prevailing role of advanced visualization is illustrated by showing several examples l ,
including ray tracing and contour plots.
1 INTRODUCTION TO EUROPEAN HYPERSONIC ACTIVITIES
In Europe, at present several hypersonic programs exist where the major activity is the
Hermes program. Besides the Hermes activities, more advanced space transportation
IThe examples in this chapter were originally produced in colour; the black and white reproductions may therefore show less detail than the originals.

448 Computational Methods in Hypersonic Aerodynamics


systems are being investigated on a national basis in Germany, the United Kingdom,
and also in France. These activities aim at a future European launcher that can
soar into Low Earth Orbit, LEO, by taking-off from standard airport runways. The
respective projects are named Sanger, after Dr. E. Sanger, who proposed such an
airplane in 1933, Hotol (Horizontal Take off and Landing), and AGV (1' Avion a
Grande Vitesse). The main companies involved are MBB/ERNO, MTU and Dornier,
now Deutsche Aerospace, in Germany, British Aerospace and Rolls Royce in the UK,
and Aerospatiale and Dassault Aviation in France. It is most important to have
the experience from Hermes, in order to dispose of the data needed for hypersonic
aerodynamics and thermodynamics for the new critical technologies. A reduction of
launch cost will be mandatory for long term space exploration and commercialization.
Regarding the overall cost of that program it is clear that it can only be realized on
a European scale. While Hermes' maiden flight is scheduled for 2000 and is planned
to be in operation by 2003, Sanger will not become operational before the year 2015.
However, a commitment must be made by the mid-nineties to meet that date. In
the mean time, the necessary technological and organizational preparations should be
made. In contrast to that, the American NASP (National Aero Space Plane), having
single-stage-to-orbit capability, is scheduled for its first research flight in 1996, but
could be delayed by several years.
ESA has also chosen the Huygens/Titan planetary probe as the agency's next
scientific project. The mission aims to explore the Saturn moon Titan, the atmosphere
of which is believed to consist mainly of nitrogen and some methane. The ESA Titan
probe - named Huygens after the Dutch astronomer who discovered Titan in 1656
- will be launched in 1996 and should reach Saturn in 2002. The orbiter will target
and release the probe. The payload consisting of various scientific instruments will
have a mass of about 192 kg. The probe will be an aeroassisted vehicle entering the
outer fringes of Titan's atmosphere at a velocity of about 7 km/sec. Aerobraking will
also be useful for reusable orbit transfer vehicles that will shuttle between LEO (Low
Earth Orbit) and GEO (Geosynchronous Earth Orbit) or the Moon for the delivery of
equipment, cargo or personnel and for future unmanned Mars missions for geophysical
research. An aerobraking vehicle similar to the Huygens/Titan probe will be used.
The structure will have a low ballistic coefficient and a hypersonic lift to drag ratio
of some 0.2 to 0.3. This will result in doubling the potential payload when compared
with systems using propulsive braking.
Although CFD (Computational Fluid Dynamics) is advancing rapidly, code validation along with the construction of computational grids is still a crucial issue. From
the computational side, fast and efficient mesh generation for complex geometries is
a major issue [5, 6, 12]. Recently, general, sophisticated grid generation codes as, for
example Eagle[7], gridgen[8], and Grid* [9], have become available.
In the last few years several studies have pointed out the importance of reusable
winged launchers and aerobraking vehicles for hypersonic flight. For winged launchers, the increase in speed, however, results in higher drag and increased thermal load

Computational Methods in Hypersonic Aerodynamics 449


for the future hypersonic airplane. This vehicle demands a high lift-drag (LID) ratio
with airbreathing propulsion [12]. The main aerodynamic requirements for a transatmospheric vehicle are minimizing drag; a high LID ratio of about 5-6; integration of
airframe and propulsion system; determination of flow field geometry resulting from
interaction of forebody; multiple inlet and wing; prediction of the 3D flowfield on the
external nozzle to determine nozzle thrust coefficient and thrust vector; incorporation
of air chemistry and chemistry of combustor products; optimization of longitudinal
control characteristics; and minimization of thermal loads. In contrast to a hypersonic
airplane, aerobraking (atmospheric braking), orbital transfer vehicles need a low LID
ratio. For the Titan probe, it is vital to know the contamination resulting from the
nose material in order to avoid any disturbances of the measurements. The aerodynamic variables therefore must be determined with high accuracy. Here radiation,
ionization and thermo-chemical nonequilibrium are essential to be incorporated in the
physical model.
2 PHYSICAL EFFECTS OF HIGH SPEED FLOWS
All equations discussed in this paper assume that the free mean path I of a gas molecule
is much smaller than a characteristic length L, which may be the length of the body,
the radius of nose curvature or the thickness 8 of the the BL (boundary layer). In
general, the assumption 1 -+ 0 is used in continuum fluid mechanics. The ratio IlL
is called the Knudsen number Kn. For Kn < 0.01 the continuum hypothesis is valid,
which is true up to a height of about 80 km. Locally, because of strong expansion in
high speed flows, Kn may be larger, leading to a so called Knudsen layer [29].
Within the continuum range four flight regimes, namely, subsonic, transonic, supersonic and hypersonic, are discerned, characterized by the Mach number Ma = vi c.
Flows with Ma > 5 are called hypersonic. However, to characterize high speed flow,
Ma is replaced by the actual speed v which is a measure for the temperature T via
the relation:
c

"fR
=-_.

with R

= 287 J/kgK

(air) and cp

"f -1'

"f

= 1.4

(air, perfect gas)

(1)

= 1004 J/kgK.

For example if v = 2000 m/s (Ma = 6.3) one has v 2 /2 ~ 2 * 106 J/kg, leading to
a tlT = 2000 K. The velocity of the Space Shuttle, flying at a height of 60 km is
approximately 5 km/s, resulting in a tlT of 12500 K for "f = 1.4.
It is, however, well known that vibrational excitation of molecules starts at 800 K
(1 km/s), dissociation of oxygen begins at 2500 K and nitrogen dissociates at 4000 K
(5 km/s). Ionization begins at 9000 K (10 km/s). Vibrational excitation and chemical
reactions as well as ionization lead to new degrees of freedom which consume energy
and thus lower the translational temperature T. Hence, a lower value of "f occurs,

450 Computational Methods in Hypersonic Aerodynamics


which is no longer a constant. In addition, the transport properties: viscosity JL and
thermal conductivity A vary with temperature.
For the Apollo return mission from the moon the radiation heating from the high
temperature shock layer was only slightly less than the convective heating. For the
planned Mars mission a reentry of about 6 km/s and for the Rosetta program, a
comet nucleus sampling mission, reentry velocities of about 15 km/s are expected.
The latter leads to a radiative dominated reentry. The equations presented in section
4 describe reactive flows in thermo-chemical non-equilibrium but do not account for
ionization and radiation from a gas. They are, however, valid for the major part of
the trajectory of the Space Shuttle, Hermes, NASP, and Sanger vehicles.
At very high speed only a thin shock layer exists, reducing the volume to which
the heat, converted from the kinetic energy of the vehicle, can be distributed. For
very high Mach numbers the relation /3/() = ~ holds where /3 is the shock wave
angle and () is the deflection angle. Entropy increases with a shock depending on the
strength, being largest along the centerline. Therefore an entropy gradient develops
which is related to vorticity by Crocco's theorem. Hence, high speed flows past blunt
bodies are vortical flows.
Viscosity increases with temperature thus thickening the BL. Under the assumption that the pressure p is constant in the normal direction to the surface, and using
the equation of state p = iT, a decrease in density is predicted. In order to provide
the mass flow rate at reduced density, the BL thickness increases. For a flat plate
the laminar value for 8 is obtained from ~ ~ ~ where Re is the local Reynolds
number at the BL edge. Hence, the region where inviscid flow is encountered is moved
outward. The interaction between inviscid flow and BL is called viscous interaction.
In the following this value of 8 will be used to calculate the Knudsen number to determine whether rarefied gas effects have to be considered. Let Re = cl
where L
J.'
may be any characteristic length and JL = ~plvm where I is the mean free path and Vrn
the mean molecular speed. This velocity is given by VM = )3';; where a Maxwellian
distribution is assumed and m is the mass of a single molecule. Using R = ?R /M and
?R = k NA as well as a 2 = 'YRT where?R is the universal gas constant, M the molecular
mass, k Boltzmann's constant and NA the Avogadro number, and inserting all these
terms in

(2)
In order to find out whether there is locally rarefied flow, L has to be replaced by the
BL thickness 8. This leads to
I
1
Kn = - ex: - - = =
(3)
8 MaJReL
for the hypersonic BL. All values for Re as well as the Ma number are calculated at the

-------

Computational Methods in Hypersonic Aerodynamics 451


edge of the BL. With a vehicle speed of 5000 m/s at a height of 60 km the following
values are taken from the standard atmosphere: Too = 247 K, Poo = 3.1 * 10- 4 kg/m 3 ,
and 1'00 = 1.58 * 10- 5 ~;. The length of the Space Shuttle is about 30 m, which results
in ReL ~ 3 * 106 From Ma oo = voo(rRTtl/2 one obtains Ma ~ 18. Thus Kn ~
0.01, which clearly shows that the flow is in the continuum range. However, it should
be noted that local areas in the BL may exist where this is invalid, since the density
in the BL may be substantially reduced, as was discussed above, forming a so-called
Knudsen layer. If such a layer exists, no slip BCs are no longer accurate, and a slip
velocity at the wall has to be used. From simple gas kinetic considerations one finds

v =/ W
d

(OV)
on

(4)
W

where d is the accomodation coefficient, denoting the fraction of molecules that are
diffusively reflected [26, 32J; d=l for continuum flow which leads to Vw i= 0 even for
this flow. However, the mean free path I of ~ 10- 7 m for p = 1 atm and T = 273
K, practically leads to Vw = O. The accomodation coefficient strongly depends on
the surface material and the gas and is a measure of the gas-surface interaction. A
reduction of d leads to a reduction in heat transfer to the wall. The mean free Piith
can be obtained by the following simple physical model. Let n be the number of
molecules per unit volume. Consider a volume of molecules of height I and unit area.
Let ro be the radius of a molecule. Now project all molecules nl on the unit area.
If the unit area is completely covered by the area of the molecules, it is opaque and
the probability for a collision is one. Since molecules collide if they are a distance 2ro
apart, one finds
nI47rT~

=1

or

1----- n411'r5 - na

(5)

where a = 411'r~ is the collision cross section. The average value for a molecule of the
collision cross section is a = 411'(10- 10 )2 ~ 1O- 19 m 2 At 273 K and 1 atm, the number
of molecules is n =2.7 * 1025 1/m3 leading to I ~ 3 * 1O- 7 m, which was used above.
The collision frequency T = vrn/I for N2 has a value of about 2 * 109 8- 1 with Vrn ~
600 m/s.
Another difference between ideal and real gas flow (high temperature effects) is
the following. For ideal gas flow with a high-intensity shock wave, that is P2 ~ PI,
the density ratio ahead and behind the shock is given by the equation:
P2
,+ 1
-=--,
PI
,-1

(6)

which leads to P2 = 4Pl for a monoatomic gas and to P2 = 6Pl for a diatomic gas.
The temperature ratio T2/Tl increases unlimited with pl/ PI, the proportionality factor
being~. Since in a real gas flow additional degrees of freedom exist, the temperature
behind the shock is much lower than for a perfect gas. Dissociation leads to a much

452 Computational Methods in Hypersonic Aerodynamics


higher ratio of pd PI in the real gas case. The pressure ratio
higher in the real gas case for velocities larger than 2 km/s.

pdPI

is only marginally

For high Ma numbers, the static pressure Poo can be neglected in comparison with
the dynamic pressure qoo and the static energy ~T can be neglected with regard to
the kinetic energy (per unit mass). In [23] an equation for the shock-standoff distance
is given:

(7)
where RN is the radius of curvature (e.g. Hermes nose radius). Since the ratio P2/ PI
is larger for a real gas flow than for a perfect gas, ~x decreases and the shock moves
closer to the body. The shock is closest for chemical equilibrium, since the highest
dissociation rate takes place in this case. For non-equilibrium flow, the shock standoff
distance is between those of an ideal gas and a gas in chemical equilibrium.
The heat transfer rate q".v, measured in J/(m 2 s), into the aerodynamic surface is
of prevailing concern for reentry vehicles. From the definition of the Stanton number
CH the following equation holds:

(8)
where PooVoo are the free stream density and velocity. The enthalpy haw denotes
the adiabatic wall enthalpy, that is the enthalpy obtained for the flow adiabatically
brought to rest at the wall. Of importance is also the so-called recovery factor r,
which denotes the fraction of kinetic energy converted into thermal energy within the
BL. Hence one can write

(9)
For laminar flow r = ffr and for turbulent flow r is Pri . The recovery temperature
Tr can therefore be written as

Tr = Too

r 1 2
+ --2voo
c

(10)

The heat transfer rate is determined by the temperature difference Tr - Tw and also
by the mechanism of conversion of kinetic energy into thermal energy by friction. The
Stanton number gives the fraction of the energy flux that will go into the aerodynamic
surface. During reentry the kinetic energy flux can reach up to 20 MW /m 2 [23] and
not more than a few tenths of a percent of this amount must go to the aerodynamic
surface. However, since Tw cannot exceed certain limits due to material constraints, q".v
strongly increases with higher flow velocity. Therefore active cooling and reradiation
are needed.
In this case the heat flux

q".v takes the form:


.

qw

1 C 3
= "2Poo
ilvoo

(11)

Computational Methods in Hypersonic Aerodynamics 453


where hw and hoc are neglected with respect to haw' according to our remarks above
that static quantities are much smaller than the dynamic values. The heat flux is
also influenced by chemistry, in particular by catalytic effects at the wall. It has been
noticed that the catalycity of the TPS (thermal protection system) of the Space Shuttle changes from flight to flight. An accurate prediction of surface temperature may
therefore be very difficult since this mechanism is not well understood. If there is a
transition from a laminar to a turbulent BL, the heat flux may increase by a factor
of 3 to 4 [23], as has been shown for a slender cone of 5.
3 HYPERSONIC AEROTHERMODYNAMICS AND COMPUTATIONAL FLUID
DYNAMICS
3.1 Aspects of Numerical Simulation of Hypersonic Flow
The design of new hypersonic vehicles will to a large extent depend on numerical
simulation, since ground based facilities cannot duplicate all aerodynamic simulation
parameters at the same time. For the Titan entry e.g. present windtunnels cannot
simulate the high enthalpy and the low atmospheric density during the aerobraking
period along with the proper chemical composition.
Over the last decade CFD has made great strides in its ability to compute external and internal flows. Great improvements have been made in the ability to predict
hypersonic external flows. The main interest in the 1970's focused on the flow prediction of reentry bodies and the space shuttle. For complete configurations such as
the space shuttle, 3D Euler codes were developed and coupled with approximate 3D
viscous codes to provide the required solutions. In addition, time-dependent Euler
codes have been developed, and in recent years parabolized Navier-Stokes, PNS, and
complete N-S codes have been used for complete aircraft and spacecraft configurations.
These calculations still cannot be done on a routine basis but demand in many
cases excessive time for input preparation, e.g. 3D grid generation, and also require
very high computation times, making CFD a costly design tool. In addition, usage
of sophisticated 3D Euler or N-S solvers demands a thorough understanding of the
underlying physics and numerical solution techniques. For example, first calculations
for the Mach 21 case lead to negative densities for the Hermes configuration but not
for the double ellipsoid. The reason for this behavior was the increased length of
Hermes leading to a larger flow expansion. In many cases, however, CFD is less costly
than the use of ground based facilities. In addition, it allows the separation of the
physical effects, and every condition that might occur in flight can be simulated in
advance.
While a flowfield analysis of a complex 3D configuration is possible, the question
of the validation of these codes is still not settled, in particular when complex physical
phenomena are present, e.g. shock interactions, transition from laminar to turbulent
flow, viscous interactions, separation etc.

454 Computational Methods in Hypersonic Aerodynamics


An important point is the design of configurations themselves, which are subject
to certain constraints. Therefore the use of CFD codes in the optimal design of hypersonic vehicles along with the proper visualization of the large amount of computed
data has to be substantially advanced. What is needed is a CFD WORKBENCH
for the aerodynamical engineer. With the advent of massively parallel computers and
new powerful graphics workstations (e.g. Silicon Graphics, Hewlett Packard, IBM
RS/6000), a rapid maturation of CFD can be foreseen [15, 16, 20, 21]. Recently, the
use of expert systems for input preparation (e.g. domain composition of the solution
area) and for the selection of numerical schemes has been investigated.
The solution method most widely used for aerodynamic configurations for complex 3D geometries including viscous effects as well as chemistry is the finite volume
method (FVM), employing curvilinear coordinates. Although sometimes finite element methods (FEM) are utilized, there are substantial advantages in using FVs to
simulate the above mentioned physical phenomena.
Comparing the numerical efficiency of the two methods, test computations for
the time-dependent SWEs (Shallow Water Equation), a type of Euler equation where
pressure is replaced by the elevation of a water column, showed that for the same
solution accuracy the computing time for the FEM was a factor of ten higher than for
the finite differences. This is understandable, since the FEM demands the inversion of
a full matrix at each time step. A discussion of structured versus unstructured grids
is given in the next section, expressing the personal view and experience of the authors.
3.2 Numerical Grid Generation: Structured versus Unstructured Grids
Regarding high speed flow past 3D objects many flow situations can be encountered
where the flow in the vicinity of the body is aligned with the surface, i.e. there is a
prevailing flow direction. All analytic inviscid solutions in 2D (ramp) or 3D (cone)
are based on the fact that behind the shock the flow is parallel to the walls. This
is especially true in the case of hypersonic flow because of the high kinetic energy.
The use of a SG (Structured Grid), also called body fitted or boundary fitted grid,
allows the alignment of the finite FVs in that direction, resulting in locally quasi 1D
flow. Hence, numerical diffusion can be reduced, i.e. better accuracy is achieved when
compared to an UG (Unstructured Grid). Second, SGs can be made orthogonal at
boundaries, facilitating the implementation of BCs (Boundary Condition) and also
increasing the numerical accuracy at boundaries. Furthermore, orthogonality will
increase the accuracy when algebraic turbulence models are employed. In the solution
ofthe N-S (Navier-Stokes) equations, the BL (Boundary Layer) must be resolved. This
demands that the grid is closely wrapped around the body to describe the physics
of the BL (some 32 layers are used in general for SGs or UGs). Here some type of
SG is indispensable. In addition, to describe the surface of the body a structured
approach is better suited. The resolution of the BL leads to large anisotropies in the
length scales in the directions along and off the body. Since the time-step size in an
explicit scheme is governed by the smallest length scale or, in the case of chemical

--

-----

Computational Methods in Hypersonic Aerodynamics 455

reacting flow, by the magnitude of the chemical production terms, extremely small
time steps will be necessary. This behavior is not demanded by accuracy but to retain
the stability of the scheme. Thus, implicit schemes will be of advantage. In order to
invert the implicit operator, factorization is generally used, resulting in two factors if
LU decomposition (that is factoring in the direction of the plus and minus eigenvalues
of the J acobians) is employed, or in three factors if the coordinate directions are
used. For the unstructured approach there is no direct way to perform this type of
factorization. Moreover, the use of the so-called thin layer approach, that is retaining
the viscous terms only in the direction off the body, reduces the computer time by
about 30%. Since there are no coordinate lines in the UG, this simplification is not
possible. A fairly complex procedure would be needed to artificially construct these
lines. Moreover, the flow solver based on the UG is substantially slower than for the
SG. This is due to the more complicated data structure needed for UGs. Factors of
3, and by some authors of up to 10, have been given in the literature.
In order to calculate heat loads for vehicles flying below Ma 8, turbulence models
have to be used, for example, the difference in surface temperature for Sanger at
cruising speed (around Ma 5) is about 500 K depending on laminar or turbulent
flow calculations. Depending on the real surface temperature encountered in flight
a totally different type of vehicle has to be designed, since a cooled Titanium wall
cannot withstand a temperature of 1300 K for a long period (about 20 minutes).
Therefore turbulent calculations are of extreme importance. Only a SG can provide
the alignment along with the orthogonality at the boundary to accurately perform
these calculations.
Although it might be thought that CPU time is no longer a critical item with the
next generation of computers, this will not be true if turbulent flows and transition
phenomena or flows in thermo-chemical non-equilibrium are to be modeled. For the
latter type of flow even 2D computations with 10 4 grid points will result in a computing
time which is in the magnitude of hours on a CRAY XMP. Suppose there is a Cray4
that is a 100 times more powerful than the present Cray2 and suppose we need a
factor of three more in CPU time and memory based on 10 Million grid points, this
would amount to an additional 299 present day Cray2s, and using 100 words per
gridpoint would demand an additional 16 Gbytes of memory. Since transition and
turbulence are the driving force for future applications in aerospace, any additional
increase in computing speed and memory has to be used to improve the solution of
these physical phenomena. However, for the coupled non-equilibrium equations an
acceleration technique which also guarantees positivity is available [18].
An important point for the accuracy of the solution is the capability of grid point
clustering and solution adaptation. In general, SGs provide sophisticated means both
for clustering and adaptation using redistribution or local enrichment techniques. A
comparison of these two approaches is given by Dannenhoffer [1] where local enrichment gives somewhat better results. However, it is much more costly to use. In many
cases of practical interest, for example fine resolution of a bow shock or a canopy shock

456 Computational Methods in Hypersonic Aerodynamics


and in situations where shocks are reflected, the alignment of the grid can result in
a more accurate solution than randomly filling the space with an enormously large
number of smaller and smaller tetrahedrons or hexahedrons. The highest degree of
freedom of course is obtained in UGs. It is therefore felt, however, that mesh redistribution and alignment is totally adequate for the major part of the flow situations
encountered in external flows, especially in aerodynamics. The large majority of physical phenomena encountered in external and also in internal flows exhibit certain well
ordered structures, such as a bow shock, or system of reflected shocks or some type of
shock-shock interaction which can be perfectly matched by adaptive grid alignment,
coupled to the solution process. Only if a very complex wave pattern evolves due to
special physical phenomena, for example, generating dozens of shock waves, the UG
has advantages. In addition, the coupling of SGs with UGs is possible as has been
shown by Shaw et al. (see AGARD Report of Ref. [8]). Such a grid is called a hybrid
grid. A mixture between the boundary fitted grid approach and the unstructured
approach is the use of multiblock grids, which is followed in Grid* [37]. On the block
level the grid is unstructured but in each block the grid is structured and slope continuity is provided across block boundaries. Recently a 94 block grid for the Space
Shuttle has been generated, demonstrating the variability of this approach.
4 MATHEMATICAL FORMULATION OF GOVERNING EQUATIONS
4.1 Governing Physical Equations
The formulation of the governing physical equations with and without chemical effects
is given in integral form to ensure conservativity. The integral can be carried out in
2D or in 3D. For the sake of simplicity only two spatial dimensions are used but the
extension to three dimensions is straightforward. This also holds true for the numerical
solution scheme presented. All equations derived in this section are valid for any of
the modern numerical solution techniques. The denotes Cartesian coordinates while
the transformed quantities are denoted by their proper symbols. The N-S equations
are given in conservative form in equation (15) where U is the vector of the primitive
variables, and F, G denote flux vectors. Vector W is the chemical source term, derived
from the formation of species i, which is governed by the respective rate equation.
The derivation starts from the integral form of the N-S equations (the Navier-Stokes
equations here include the chemistry) [15, 24, 25]

aat Jv[ U dV +

A(V)

FA

= [ W dV

Subdividing the SD (Solution Domain) into a set of elemental volumes


formulation
U
aa ~V +
FdA = L:W~V

L:
~v

L:
[
~v JA(V)

(12)

Jv

~v

~V

yields the
(13)

where U is the vector of conserved variables, averaged over an arbitrary elemental


volume ~V. A(~V) denotes the closed surface of volume ~V, F is the flux tensor

Computational Methods in Hypersonic Aerodynamics 457


and W is the source term arising from the chemical reactions. It should be noted that
the above formulation is valid for any arbitrary CS (Coordinate System). A different
chemical model will result in a different number of species and vibrational energies,
but the character of the resulting system of PDEs will remain unchanged, in that
all the quantities are described by transport equations. This also remains valid if a
transport model for turbulence is added, e.g. a two equation model for k and for k
and w. It is interesting to note that a turbulence model of that type exhibits similar
numerical properties as a chemistry model stiffness and the requirement of positivity.
Therefore, the solution techniques for chemistry can be applied. The species densities
and vibrational energies have been arranged together.

(14)

(15)
The first five entries in U indicate the species N 2 , O2 , NO, N, and 0, followed by the
two momentum components, and the three vibrational temperatures for N 2 , O 2 , and
NO. The last variable, E[J/m3], is the total energy for the translational-rotational
modes. The equation for E is obtained by summing over all species energies, Ei and
by introducing a bulk velocity u, density p, and pressure p, which are defined as
follows:
5

p.

p.

E:= LEiiP:= LPiiP:= L ~.RTiU:= L "":"UiiUi := Ui - u.


i=1
i=1
i=1'
i=1 P

(16)

The translational-rotational temperature, T, is calculated from the equation


E

i=1

i=1

= LPi(cviT + hn + LEvi + 1/2p(u 2 + v 2 ).

(17)

For temperatures below 600 D C, no chemical effects occur, and subsequently neither
vibrational energies nor additional chemical species exist. There is only one density
and pressure, and the expression for the translational energy has to be modified accordingly. Source vector W is identically zero. The quantity Ui is the diffusion velocity
in x direction of species i. The first 5 components of Ware the species production
terms having dimension [kg/m 3 sJ. The rate equation for species i, which is of dimension mole/m 3 s, has therefore to be multiplied by molecular weight, M;[kg/mole],
to result in the correct dimension. As the momentum terms are not influenced by
species production, the next two entries are O. The terms, Evi, (i = 1,2,3), are determined from the Landau-Teller relation. In this formulation a vibrational-vibrational
coupling is not included, which may lead to equalization of vibrational temperatures.
Expressions, eviwi, account for the production of species N 2 , O2 , and NO. Quantities
evi have the dimension [J/kg]. W has the form:

458 Computational Methods in Hypersonic Aerodynamics


(18)
Expressions e~wt account for the production of vibrational energies due to the production of the respective species. The terms e~w; describe the decrease of vibrational
energy due to consumption of this species. In equation (18) flux vector, F, is given.
G is obtained in a similar way. The species velocity Ui is written in two parts, namely
the bulk velocity u, and the diffusion velocity Ui, see equation (16). According to this
formulation, the term E~=l PihiUi arises where hi is the enthalpy of species i per unit
volume, that is hi = Ei + Pi. Heat flux terms, qui" are included in the vibrational
energy balance, see below. The component of the flux vector in x direction, F, has
the form:
PI(U + Ud
P2(U + ( 2)
P3(U + ( 3)
P4(U + ( 4)
Ps(u + Us)
Evl (u + UI) - qvl"
Ev2 (u + ( 2) - qv2"
E v3 (U + ( 3) - qv3"
2
pu + P - T""
pUV - T"y
(E + p)u - T""U - T"yV - q" -

,G= ...

(19)

E pahsUa

where subscripts 1 to 5 denote the species N 2 , O2 , NO, N, and O. To close the system,
the equation for the total pressure is determined from the partial pressures using the
equation of state for a calorically perfect gas for each species (see equation 16). The
fluxes can be separated into their inviscid and viscous parts, as shown in equation

(20)
(20)

FI =

PIU
P2 U
P3 U
P4 U
PSU
Evlu
E v2 u
E v3 u
pu 2 +p
pUV
(E + p)u

Fv=

PIUI
P2U2
P3U3
P4U4
psUs
Evl UI - qvl"
E v2 U2 - qv2"
E v3 U3 - qv3"
-T""
-T"y
-T""U - T"yV - q" - E PahsUs
(21 )

Computational Methods in Hypersonic Aerodynamics 459

A similar relation holds for Gland G v . For inviscid flow all viscous fluxes are
zero. In case there are no chemical reactions, only one species is present, denoted by
p and the Euler or the Navier-Stokes equations are obtained respectively.
4.2 Discretized Form of Physical Equations

Since the governing physical equations are solved on the regular geometry of a block
in the CP, the following transformation has to be used (2D, similar in 3D):

e= e(x,y)
x
77

= X(e,77)
= 77(X,y)

(22)

y = y(e, 77)
The equations are transformed starting from the integral form of Eq.(13). To carry
out the integration, only grid point positions are needed. The rest of the geometrical
information, such as volume, area, surface vectors etc., is contained in the metric
coefficients, which can be numerically calculated from the specified grid point distributions, and are provided by the grid generation program. Along with the governing
physical equations, the BCs have to be transformed. Using the BFG approach converts the solution of a set of PDEs on a complex domain to the solution of a system of
PDEs of the same type on a rectangle (2D) or box (3D), introducing some additional
terms. In the case of a block-structured grid, a set of connected rectangles or boxes
are used. Carrying out the integration over an elemental volume yields in 2D

a(UJ);.1
at

A]
[A
A]
+ [A
J(Fex + Ge i+l/2,i - J(Fex + Ge i-l/2,j
y)

y)

(23)

+ [J(F77x + G77y)L,i+I/2 -

[J(F77x

+ G77y)L.i_l/2 = (W J)i,i

where half point values denote proper cell surfaces. From now on U denotes the
product (; J and F, G are used for the contravariant fluxes, i.e. the fluxes in the 77
directions. The relation between the transformed and Cartesian fluxes for a stationary
SD is

e,

F:= J (Fex + Gey) ;


G:= J (F77x + G77Y)

(24)

where J = (ex77y - ey77x)-l is the Jacobian of the transformation. A large body of


literature exists for the numerical solution of the discretized equations, Eqs. (23). The
solutions in this paper have been obtained by some type of flux correcting transport,
combining an oscillation free first order upwind technique with a higher order technique, based on characteristic extrapolation [14, 20, 21].

460 Computational Methods in Hypersonic Aerodynamics


5 THE CHEMICAL SOURCE VECTOR W
5.1 Chemical Relations
For reentry or entry trajectories high temperature effects have to be accounted for.
The chemical source terms, Wi, for the five chemical species, used here namely N 2 , O 2 ,
NO, N, and 0 are derived from the dynamics of the chemical reactions. The main
chemical reactions to be considered for hypersonic flow, without ionization, are

N2+M
02+ M
NO+M
N 2 +O
NO+O

-t
-t
-t
-t
-t

2N+M
20+M
N+O+M
NO+N
02+ N

(25)

where M represents any of the five species, see Candler [15]. The velocity by which
the reactions proceed is dependent on the collision frequency of the molecules in the
mixture and the probability that the available energy in the collision will initiate the
reaction. The collision frequency for a particular reaction is proportional to the product of the number of molecules per unit volume ( [p;j M i ] ) of each of the reactants, of
the reaction, and the average velocity of the molecules (translational temperature).
The probability that the available energy in the reaction will initiate the reaction is
a function of the translational and vibrational energies of the reactants. The average velocity of the reactants for the collision frequency is modeled by the Arrhenius
equation. Parameter k is called the forward rate coefficient for forward reactions or
the backward rate coefficient for the reversed reactions. The Arrhenius equation is a
curve fit obtained from experimental data

(26)
where C is called the frequency factor, 7J is needed for improved data fitting and TA is
the activation temperature of the reaction. The reaction temperature Tr is dependent
on the nature of the reaction, it represents the amount of energy available in a collision.
In most cases the reaction temperature is equal to the translational temperature,
however, when, for example, a diatomic molecule collides with a third body and
dissociates, the reaction temperature is taken to be the geometrical average of the
diatomic vibrational temperature and of the translational temperature, as suggested
by Park [31].

(27)
In a later paper a value of Tr = TOIT;;-OI with a = 0.7 is used. Table 1 lists the
coefficients for the Arrhenius equation for all major reactions in air along with the
respective reaction temperatures. The reaction rates for each of the reactions is determined by the species densities and the forward and backward reaction rate coefficients.
Since the forward and backward reactions occur simultaneously, the complete reaction

Computational Methods in Hypersonic Aerodynamics 461

reaction
N2 + N2

4.7 * 10
2.72 * 10 10
N2 + O2 +- 2N + O2
1.9 * 1014
--+
1.1 * 1010
N2 + NO +- 2N + NO
1.9 * 10 14
--+
1.1 * 10 10
N2 +N +- 2N +N
4.09 * 1019
--+
2.27 * 10 15
N2 + 0 +- 2N + 0
1.9 * 1014
--+
1.1 * 10 10
O2 + N2 +- 20 + N2
7.2 * 10 15
--+
6.0 * 109
O2 + O2 +- 20 + O 2
3.24 * 10 16
--+
2.7*10 10
O 2 + NO +- 20 + NO
3.6 * 10 15
--+
3.0 * 109
O2 +N +- 20 +N
3.6 * 10 15
--+
3.0 * 109
O2 + 0 +- 20 + 0
9.0 * 1016
--+
7.5 * 1010
3.9 * 10 17
NO + N2 +- N + 0 + N2
--+
1.0 * 10 14
NO + O2 +- N + 0 + O2
3.9 * 10 17
--+
1.0 * 1014
NO+NO+-N+O+NO 7.8 * 1018
--+
2.0 * 1015
NO + N +- N + 0 + N
7.8 * 10 18
--+
2.0 * 10 15
NO+O+-N+O+O
7.8 * 1018
--+
2.0 * 1015
N2 +0 +- NO+N
7.0 * 10 10
--+
1.56 * 1010
NO+O +- O 2 +N
3.2 * 106
--+
1.3 * 107
+-

2N + N2

14

--+

Table 1: Arrhenius coefficients

1]

TA[OK]

-0.5
-0.5
-0.5
-0.5
-0.5
-0.5
-1.5
-1.5
-0.5
-0.5
-1.0
-0.5
-1.0
-0.5
-1.0
-0.5
-1.0
-0.5
-1.0
-0.5
-1.5
-1.5
-1.5
-1.5
-1.5
-1.5
-1.5
-1.5
-1.5
-1.5

113000
0
113000
0
113000
0
113000
0
113000
0
59500
0
59500
0
59500
0
59500
0
59500
0
75500
0
75500
0
75500
0
75500
0
75500
0
38000
0
19700
3580

0
0.0
1.0
1.0

[kmo7'3 sec

or

Tr
JTvN2 T
T
JTvN2 T
T
JTvN2 T
T
JTvN2 T
T
JTvN2 T
T
[Tv0 2T
T
JTV02 T
T
/TV02T
T
JTV02 T
T
[Tv0 2T
T
,.fTvNOT
T
TvNOT
T
TvNOT
T
VTvNO
T
,.fTvNO
T
/.TVN2T
vlTvNOT
TVNOT

kmol",;,6 sec]

Note: The coefficient C is very sensitive to the 1] value selected. Therefore, special
care must be taken when these coefficients are compared with those from other sources.

462 Computational Methods in Hypersonic Aerodynamics

[mole/m 3 s] rates are given by equations (27),

R1 -- "'s
[k 11m .P.l....m..
L...m=1
Ml Mm

kblm .J.....J.....m..]
M. M. Mm

R3 --

kb3m .J....
A.m..]
M. M~ Mm

",S
L...m=1

[k 13m .J&..m..
M3 Mm

(28)

where variables kl and kb can be directly determined from Eqs.(26). During the
computations performed for the equilibrium nozzle flow (section 8.4) negative densities
occurred at the beginning of the calculations depending on the magnitude of the CFLnumber, despite the fact that a fully implicit and conservative upwind scheme was
used. An investigation revealed the source of this behavior to be the linearization
of the fluxes, Eqs.(26-32). Employing a semi-implicit discretization to the RHS of
Eqs.(15), results in a positive and stable scheme as shown in [33].
The reaction rates give the magnitude by which the reactants are consumed and
by which the products are generated. The source term Wi is obtained by grouping
together all the reaction rates that correspond to the production (+) or the consumption (-) of species i. The multiplication with Mdkg/mole] is necessary to obtain the
correct dimension. Source terms W/ and Wi- are formed by collecting terms with
positive and negative signs, respectively.

Wl=
W2 =
W3 =
W4 =
Ws =

Ml
M2
M3
M4

[-Rl - R 4]
[-R2 + Rs]
[-R3 + R4 - Rs]
[2Rl + R3 + R4 + Rs]
Ms [2R2 + R3 - R4 - Rs]

(29)

The main influence of the reaction is its release or absorption of energy. The heat
formation hi quantifies this amount of energy. For example the A(h'A) + B(hB) --+
C( he) reaction releases h'A + hB - he Joule/reaction. The two reference species in
our model are Nand 0, consequently their heat formation is set to 0 [Joule/kg]. The
heat of formation of N 2 , O2 are -33786, -15577, -21056 kJ/kg respectively [4]. The
heat of formation enters in the mathematical description of the flowfield through the
energy Eq.(16).
5.2 Thermal Relations
The molecular energy consists of five energy modes, i.e., translational, rotational,
vibrational, electronical and heat of formation (see 7.1). For diatomic molecules, the

Computational Methods in Hypersonic Aerodynamics 463


molecular energy is given by:
e

= ~ RT + RTr + ~ ()
2

eT. - 1

+ eel + hi

(30)

with T, Tr and Tv respectively the translational, rotational and vibrational temperatures, and eel, hi denoting the electronical energy and heat of formation. For applications in the continuum regime below 100000 K the following simplifications can be
made:

Tr = Tj For all the diatomic molecules important in air it takes only about five
collisions to bring the rotational temperature in equilibrium with the translational temperature. The corresponding timescale is therefore several orders of
magnitude smaller than the chemical timescale, and thus for our applications
rotational temperature and translational temperature are assumed to have the
same value.

= 0 j The change of the electronical energy as a function of temperature is


negligible below 100000 K for two reasons: either electronic excitation starts at
extremely high temperatures (N, N 2 ) or the species are sparse at the conditions
when the electronic excitation is significant (0, O2 , NO). Since only the relative
value of energies is significant, eel is set to o.
eel

Hence equation (30) takes the form for a single species


5

ei

=-2M,.T+

()i
9

eit-1

hO
i

(31 )

The vibrational energy of species i per volume is

Evi

= Pi

().
9

'

eit -

(32)

The characteristic vibrational temperature 8 i of the diatomic species N 2 , O2 , and NO


is 3390,2278 and 2745 0 K, respectively.
The time rate of change of vibrational energy of a diatomic species is controlled
by the energy exchange in collisions with other molecules, which is mainly an energy
exchange between the translational energy and the vibrational energy. The derivation of this vibrational-translational rate equation is given in Anderson [30J. The
vibrational-translational rate equation used has the form
(33)

464 Computational Methods in Hypersonic Aerodynamics

Emeq

= Pi

Oi M
R
6
e-l i

(T : transIatlOna
'
I temperature )

(34)

As was mentioned already, W+ is the production part and lVt denotes the consumption of a species. Production of species i occurs at equilibrium energy while
consumption occurs at the current vibrational energy level. The Landau-Teller relaxation time T is a function of pressure and temperature, and is presented as a
semi-empirical equation which is valid for a temperature range from 300 to 9000 J{,
see Millikan and White [17].
101325 La ftexp [Ai
Ti

(T-i - 0.015I't) - 18.42]


-b

(35)

The coefficients Ai have the values 220, 129, and 168 for N2 , O2 , and NO respectively
MiM
an d I'ij = ~.
In a chemically reacting flow there is production and consumption of species, which
will reduce or increase the vibrational energy per unit volume, even in chemical equilibrium. This rate is proportional to the production rate of species i, Wi, and a specific
source vibrational energy of species i, evi. When the production rate of species i is
negative, the specific source vibrational energy is equal to the specific vibrational
energy of species i. With a positive production rate, the specific source vibrational
energy is related to the translational energy and the vibrational energies of reactants.
5.3 Heat of Formation
The primary difference between hypersonic and supersonic flow is the fact that in
hypersonic flow the chemical composition of the fluid is changing. The corresponding
chemical reactions consume or produce vast amounts of energy and so significantly
influence the physical properties of the fluid (T,p,p). The heat of formation hi of a
molecule is the energy produced or consumed by its formation from primary molecules.
Nand 0 are chosen to be the primary molecules, consequently their heat of formation
is O. The heat offormation for N 2,02 and NO are -33786, -15577 and -21056 kJ/kg,
respectively. The negative sign indicates that their formation is consuming energy.
5.4 Viscosity
The laminar viscosity of the gas mixture is a function of the species molar concentration and the temperature. The mathematical model consists of two parts. The first
part determines the viscosity of each species i, using the Blottner Viscosity Model
which is a curve fit of experimental data as function of temperature, Blottner [18].

Blottner Viscosity Model:


I'i

= O.lexp [(AlnT + Bi)lnT + Gi ] ,

--_._-_.

kg
msec

(36)

Computational Methods in Hypersonic Aerodynamics 465


Species
N2

02
NO
N

A-
0.0268142
0.0449290
0.0464378
0.0115572
0.0203144

B-
0.3177838
-0.0826158
-0.0335511
0.6031679
0.4294404

-11.3155513
-9.2019475
-9.5767430
-12.4327495
-11.6031403

Table 2: Blottner's coefficients


The second part determines the viscosity of the mixture as function of the species
viscosities and molar concentrations, using Wilke's Mixture Rule [19]. The complete
model is appropriate for temperatures up to 10,000 K.

.
_Pi M)
(
iXi is l
mo ar concentratIOn,
Mi P

(37)

(38)
5.5 Diffusion Velocity
It is assumed that the diffusion velocity is determined only by the gradient of mass
concentration. Pressure and temperature influences are neglected.

Ui = !!....D/Xi
Pi

ax

(39)

The diffusion coefficient is given by:


(40)
where Ci is the mass concentration, and D is derived from the expression D =
where the Schmidt number is a constant, Sc = 0.5 .

pTc,

6 SUPERCOMPUTERS AND PARALLEL COMPUTERS IN AEROSPACE


The first experience with the (by far not optimized (see [3])) code for the hypersonic
equations of section 4 for the 2D case gives a crude estimate of some 5 * 10- 5 sec
per iteration and per grid point, based on a computational power of 100 Mflops.
This number is supported by Park [31]. According to Park some 2500-5000 iteration
steps have to be performed to reach the converged state. Using a modest number of
50000 grid points and assuming 3000 iteration steps, 7500 sec would be needed. Since
an IBM 3090 VF-200 supercomputer (two vector processors) has a sustained rate of

466 Computational Methods in Hypersonic Aerodynamics


about 30 Mflops, a computational time of some 10 hours will result. This is valid
for the 2D laminar case. For a more realistic 3D problem with 1 million grid points,
which is still not sufficient for a detailed engineering analysis, some 200 hours will be
necessary, provided the convergence speed in the 3D case is not reduced, which, most
likely, will be the case. For accurate design purposes, the number of grid points can
reach 10 million for a complete vehicle.
Hypersonic equations, which include turbulence as well as thermo-chemical nonequilibrium, may also require up to 107 grid points for realistic modeling. This will
lead to 1015 or even to 1016 floating point operations to solve the problem. Clearly,
this tremendous number of floating point operations cannot be handled by a single
processor in a time and cost effective manner. The solution therefore lies in the use
of a computer architecture employing a set of processors working in parallel.
Using the overlapping multi-block approach, it is natural to ascribe one or more
blocks to each processor on which to solve the physical equations. This process is called
DC (Domain Decomposition). Multiblock grid generation is a good example for DC,
since exactly the same structure is used in the flow solver, which is computationally
much more expensive. However, it is important to parallelize the code in such a way
that there are no dependencies on the shape of the SD, i.e. in some recent papers
the SD was assumed to be of rectangular shape, which was then subdivided. For the
general case, the SD is a set of connected rectangles or boxes, and its shape must not
influence the speedup of the parallel algorithm.
Hardware performance has been increasing at a very rapid pace. Cray is working
on a 2ns gallium-arsenide based design, the Cray 3, and N EC has already shipped
its 2.9ns SX - 3. All these machines use the parallel concept. The number of
processors is small, i.e. 4-16. A fully expanded SX - 3 would be capable of some 22
Gflops in theory. The supercomputers mentioned so far are shared memory machines,
which limits the number of processors because of shared memory access conflicts.
Massively parallel systems, e.g. Intel's hypercube which is a private memory (2-64
MB), message passing system where each processor has a power of several Mflops)
can sustain several hundred or thousand processors. The iPSC /2 system from Intel,
is based on 80386 technology having a vector processor of up to several Mflops (32
Bit), a maximal private memory of 16 Mbyte and a maximum of 128 nodes. The Intel
iPSC/860 is based on the i860 processor with a peak rate of 80 MFlops (32 bit) or
60 MFlops (64 bits). A machine with 128 nodes has been installed at NASA Ames
Research Center and some applications are achieving a sustained speed of 1 GFlop.
In May 1991, a system comprising 520 i860 nodes has been installed at the California
Institute of Technology, delivering a peak rate of 32 GFlops. A project, sponsored by
DARPA, is under development where two thousand nodes are being planned to be
connected, resulting in a peak performance of 128 Gflops. Based on our experience
with a 32 processor iPSC/2 and with the Intel Touchstone machines [27], it seems
very likely that several hundred processors can be used without loss of performance
(see below). In addition, numerous fluid dynamics applications are being developed

--~--~

---

Computational Methods in Hypersonic Aerodynamics 467


for parallel machines including libraries to facilitate the parallelization process. At
United Technology, the CM-2 of Thinking Machines, a SIMD computer, is heavily
used for CFD applications and sustained rates of 1 GFlops are reported.
While parallelization on shared memory machines does not demand major changes
of the computer code, message passing systems make a certain rewriting of the code
necessary. However, the potential of message passing systems seems to be much larger
[13,22].

It is often stated that scientific programs have several per cent of serial computational work, s, that limit the speedup, a, of parallel machines to an asymptotic value
of l/s, according to Amdahl's law where s + p = 1 (normalized) and n is the number
of processors:
s+p
1
a(41)
- s + pin - s + pin .
This law is based on the question: Given the computation time on the serial computer
how long does it take on the parallel system? However, the question can also be posed
in another way: Let s', p' be the serial and parallel time spent on the parallel system
then s' + p'n is the time spent on a uniprocessor system. This gives an alternative to
Amdahl's law and results in the speedup which is more relevant in practice:
a

= s' + p'n = n s' + p'

(n - l)s'

(42)

It should be noted that DC does not demand the parallelization of the solution algorithm but is based on the partitioning of the SDj i.e. the same algorithm on different
data is executed. In that respect, the serial parts s or s' can be set to for DC and
both formulas give the same result. The important factor is the ratio reT (see below), which is a measure for the communication overhead. In general, if the solution
algorithm is parallelized, Amdahl's law gives a severe limitation of the speedup, since
for n -+ 00, a equals 1/s. If, for example, s is two percent and n is 1000, the highest
possible speedup from Amdahl's law is 50. However, this law does not account for the
fact that sand p are functions of n. As described below the number of processors, the
processor speed and the memory are not independent variables, which simply means,
if we connect more and faster processors, a larger memory is needed, leading to a
larger problem size, and thus reducing the serial part. Therefore speedup increases.
If s' equals two percent and n = 1024, the scaled sized law will give a speedup of 980,
which actually has been achieved in practice [10, 11]. However, one has to keep in
mind that sand s' are different variables. If s' denoted the serial part on a parallel
processor in floating point operations, it is not correct to set s = s'n since the solution
algorithms on the uniprocessor and parallel system are different in general.

For practical applications the type of parallel systems should be selected by the
problem that has to be solved. For example, for routine applications to compute
the flow around a spacecraft on a grid of 107 grid points, needing some 10 15 floating
point operations, computation time should be about 2 to 3 hours. Systems of 1000

468 Computational Methods in Hypersonic Aerodynamics


processors can be handled, so each processor has to do about 1011 computations, and
therefore a power of 100 MFlops per processor is needed. Assuming that 100 words,
8 bytes/word, are needed per grid point, the total memory amounts to 8 GB, that
means 8 MB of private memory for each processor, resulting in 22 grid points in
each coordinate direction. The total amount of processing time per block consists of
computation and communication time:

(43)
where we assumed that 8000 floating point operations per grid point are needed,
and 10 variables of 8 byte length per boundary point have to be communicated.
Variables t e , tT are the time per floating point operation and the transfer time per
byte, respectively. For a crude estimate, we omit the set up time for a message. Using
a bus speed of 250 MB/s (quite high, the present iPSC/2 has a band width of 2.8
MB/s for messages longer than 16 Kb), we find for the ratio of computation time and
communication time.
N 3 * 8000 * 250
(44)
reT := 6N2 * 10 * 8 * 100 r::::: 40N
That is for N = 22, communication time per block is less than 0.25 % of the computation time. In that respect, implicit schemes should be favored, because the amount
of computation per time step is much larger than for an explicit one.
In order to achieve the high computational power per node a MIMD /SIMD (M ultipIe Instruction Multiple Data; Simple Instruction Multiple Data) architecture should
be chosen; that means, the system has a massively parallel architecture, e.g. Intel
or Suprenum, and each processor itself is equipped with a pipelined floating point
processor. It should be noted that even if reT> > 1 this is not sufficient, since, if the
computation speed of the single processor is small, e.g. 0.1 MFlops, this will lead to
a large speedup which is, of course, somewhat misleading because the high value for
reT only results from the low processor performance. In conclusion, it is believed that
the concept of MIMD is most promising for computationally intensive applications in
fluid physics, and that DC will be a powerful concept to tackle problems demanding
excessive number crunching.
7 Grid* : A GENERAL GRID GENERATION TOOLBOX
7.1 Grid* Description
Grid* is a universal grid generation package for multiblock 2D, surface, and 3D
solution domains that may be of arbitrary complex geometrical shape. Its name
is derived from using the star as a wildcard which stands for the collection of the
modules that make up Grid* . The package comprises interactive tools for surface
grid generation and for grid enrichment, i.e. there are very efficient means for grid
point clustering, e.g., to obtain a Navier-Stokes grid from an Euler grid. For the
Hermes grid generation an Euler grid of 300,000 points was converted into a NavierStokes grid of 600,000 points in a few minutes on a PC. Grid* can be used in every area

Computational Methods in Hypersonic Aerodynamics 469


where structured grids are needed, and provides special features, such as outer surface
generation and grid generation on such a surface, describing the outer boundary of
the solution domain (SD) for external flow past an aircraft or spacecraft. In addition,
special projection techniques using the grid point distribution on the body surface
are available to generate the corresponding grid on the outer boundary, a nontrivial
task if the body surface has a different structure than the outer boundary. A simple
example is the Hermes Space Plane with a hyperbolic-elliptic (cross section) surface
as the outer boundary.
Grid* is a collection of C routines, fairly small in size, e.g. the 3D multiblock
grid generation module, grd3d, consists of about 1500 lines in C. The modules can be
used in arbitrary order, each designed for a special task.
The grid generation modules grd2d, grdsrf, and grd3d use a combination of algebraic and elliptic grid generation methods, where elliptic techniques are mainly used
for smoothing purposes. The grid visualization module Xivis is based on Xll, allowing this system to be run on virtually any type of computer, from a PC under SCQ
Unix, IBM 6000, Silicon Graphics to an IBM 3090 etc. All grids presented in this
paper were generated on a PC with 8 MB internal memory running SCQ Unix. To
speed up convergence, 3D blocks can be interactively cut out of the SD. Grid point
distribution on the block faces is frozen, while points in the interior are iterated. This
technique was successfully used when Grid* generated the 300,000 points Euler grid
for the Hermes Space Plane: 10 iterations with SQR were used for the complete spacecraft, while a block of about 8,000 points was cut out around the winglets and iterated
100 times to improve grid quality and then was inserted back. In Grid* dynamic storage allocation is used, so the user never has to specify any array dimension. Special
data types are used, which allow a much more compact programming. A menu allows
the selection of main modules, and a submenu is displayed after selection. The screen
is subdivided in a viewport, textport, messageport, and menu as well as a submenu
area. If a submenu is in use, the commands belonging to this submenu are displayed
instead of the menu as well as the usage of the mouse. The viewport can be increased
to a full screen. Rotation and zoom features are also available as well as a user defined
colormap. Images may be stored either as a pixelfile or in the Postscript format and
can be provided with a title. All graphics functions are running under X-Windows so
the code is fully portable.
Grid input consists of a control file, describing the connectivity and orientation
with respect to neighboring blocks as well as the control functions. Each block has its
own local coordinate system. Although the user need not specify any overlap of block
faces, all grids are slope continuous across block boundaries. Hence block boundaries
are not visible. A simple, but powerful meta language is used in the control files and
the coordinates files. Files can be in binary or may be formatted. The widely used
Plot 3d (NASA Ames) grid file format is also supported. A coupling to the NavierStokes solver exists so that the multiblock grid can be directly interfaced to the solver
without any user interaction.

470 Computational Methods in Hypersonic Aerodynamics


7.2 Grid* Modules
The Grid* package comprises several modules to generate and visualize 2D and 3D
multiblock grids. These modules are selected through a mouse driven menu, but can
also be called from the command line as stand-alone modules. All modules make use
of features of the C programming language, such as dynamic memory allocation and
user defined data structures, so that the code is both efficient and compact. Below,
the features of the various modules will be explained briefly.
Modules grd2d, grdsrf, and grd3d are used to generate 2D grids, surface grids,
and volume grids on arbitrarily shaped bodies. These modules use a combination of
algebraic and elliptic grid generation where the algebraic part is used for initialization,
and the elliptic technique has a smoothing purpose. Slope continuity is guaranteed
throughout the whole grid, including block boundaries. Special features are available
to achieve the desired grid line distribution via distance and angle control.
The grdsrf module can be used to change the grid topology and also the grid line
distribution of an existing surface grid. It takes a description of the surface as input,
normally the output of a CAD/CAM program, to determine the metric coefficients
and then redistributes the grid points on the surface, according to the blocking given
in the surface grid control file.
Module plane has been developed to interactively cut out and insert parts of the
solution domain and also to insert planes or subblocks in the SD. This module is
also used to construct the blocking of the surface grid, a process which precedes the
volume grid generation. Very often the simple technique of distributing grid points
along cross sections leads to a non-optimal grid. One can also use this module to
cut out parts of the SD which need more iterations to obtain the desired grid line
distribution. These parts can then be iterated separately and can be put back in the
SD by plane. As mentioned above, this method was used for the 300,000 point Euler
grid of the Hermes Space Plane. It was also used to construct a 4 block Shuttle grid
both for Euler and NS calculations. A 94 block Euler grid for the Space Shuttle has
been constructed incorporating the body flap. The advantage of such a multi block
grid is the geometric flexibility and the nearly equidistant and orthogonal grid. In
addition, the multi block grid leads to a substantial reduction of grid points and to a
singularity free grid, which is aligned to the flow and can be adapted to areas where
large gradients exist.
A brief description of the modules is presented below:
plane: manipulation of surface patches. Cut and paste patches to get the desired
block structure .
grd2d: multi-block 2D grid generation: algebraic initialization and elliptic
smoothing with control functions.

Computational Methods in Hypersonic Aerodynamics 471


b-ed: editor like software to build command files for large multi-block grids.
grdsrf: multi-block surface grid generator with control functions. Needs a surface
description file to determine metric coefficients.
grd3d: multi-block volume grid generator with control functions.
enrich: adaptation of grid points along with enrichment or deletion of grid lines,
planes or entire blocks.
adapt: adaptive clustering for 2D and 3D multiblock grids where weight functions are determined from the physical solution (test version only).
Xivis: 2D and 3D multi block grid visualization program running under XWindows, using the Motif user interface.
connect: construct point to point connection for neighboring blocks. Thus, a
multiblock flow solver does not need to know about the grid topology.
8 RESULTS FOR HYPERSONIC VEHICLES
In this section results will be shown for three generic shapes and for two complex 3D
vehicles. The generic test cases are an axisymmetric nozzle, a 2D ramp, and the double
ellipsoid. The vehicles presented are the Hermes Space Plane and the Shuttle Orbiter.
8.1 Grid Generation for 3D Vehicles
Grid generation is still the most time consuming process. Figure 1 shows a multi
block Navier-Stokes airfoil grid and figure 2 a 3D one for the double ellipsoid. A
grid generated by Grid* is read in by the Navier-Stokes solver, without any user
interaction. The additional overlap of 2 faces for neighboring blocks is generated
automatically. The user only has to specify BCs on the fixed surfaces of the SD. In
principle, if the grid has the quality to represent the flow pyhsics, only the solution
has to be visualized. In practice, it may happen that the solution blows up because
of instability or, for example, a negative temperature or density value is encountered,
because of the numerical approximations. Even if a Mach 25 calculation for the
double ellipsoid is successful it cannot be concluded that a Hermes computation for
the same Mach can be done. Since Hermes is a longer vehicle, expansion on the leeside
is stronger, which may eventually lead to negative values. Although one can try to
reco-ver from such a situation, it is better to avoid unphysical behavior in the solution
process (sec. 8.6). Monoblock grids are relatively straightforward to generate for Euler
solutions for a vehicle like Hermes, as shown in figures 3 and 4. The Space Shuttle is
a somewhat more difficult geometry as will be outlined below. A monoblock grid has
an O-type topology in the meridional plane and is of C-type in the streamwise plane.
However, at the nose a singularity exists since the CS is locally spherical. This leads to
finite volumes along the singular line which are not hexahedral in shape because one

472

Computational Methods in Hypersonic Aerodynamics

face shrinks to an edge. Although a finite volume technique can handle any volume
shape, there is no information transfer across the singularity. The flow information
is transported in the circumferential direction only, which leads to a slowdown of the
convergence rate, as has actually been observed in viscous Shuttle calculations [34].
Again, special measures can be taken in the flow solver to handle this case efficiently,
but, of course, the best solution is to avoid this kind of singularity. A 4 block grid has
therefore to be constructed for both the Hermes and the Space Shuttle and a much
improved convergence rate and a much better shock resolution has been obtained
(Secs. 8.6, 8.7). All grids are slope continuous. A 4 block grid, however, apart from
the nose region has the same topology as the monoblock grid. If N-S computations
have to be performed, the BL has to be resolved. This demands the surface geometry
of a vehicle to be described in sufficient detail, to avoid intersection of grid lines with
the surface. If the skin friction and the heat transfer coefficient are to be calculated,
the first grid point should be about 10 - 20 fL m from the surface. If a BL or a PNS
code is used this distance has to be reduced because of the weaker coupling of the
physical equations. An aspect ratio of 10 4 for the streamwise and the radial size of a
volume may be produced.
Figure 5 shows a Hermes 7 block grid and figure 6 its logical structure. The
numbered surfaces depict the outer surface wrapping around the vehicle and the black
lines describe the connectivity of the blocks. Block 7 denotes the nose block. Apart
from the nose block, the grid is of O-type typology in the meridional plane and of
C-type topology in the streamwise plane.
The topology of this grid is chosen such that the space between winglet and fuselage
is gridded like a Cartesian CS to optimize grid uniformity and grid orthogonality. The
N-S grid is obtained by wrapping blocks around the surface in an O-type fashion to
best resolve the BL (boundary layer) and grid orthogonality at the surface. An Euler
grid is always constructed first, and after that module "enrich" of Grid* is used to
interactively add the N-S points. In that process, it is also possible to delete grid lines
or grid planes.
An example is shown in fig. 7. Fig. 8 shows a somewhat simplified shuttle geometry without bodyflap and with part of the vertical fin discarded. In addition the final
cross section has been modeled as a 2D flat plane. A more advanced Space Shuttle
including body flap and vertical fin can be found in [7]. The shuttle is shown in figs. 9
and 10 in the display window along the data input window. The user can manipulate
colortable and dispose over four action windows which are used for file selection and
manipulation of 3D planes. A detailed description is given in [6, 7]. Finally, fig. 11
gives an example of a solution adaptive grid on a cone.
8.2 Visualization for 3D Data
In order to analyze an enormous amount of data in a synthetic way it is mandatory
to have powerful graphical tools. Fig. 12 shows an example of the streamline tracing
option of an advanced graphic package. Streamlines determined by an Euler solution

Computational Methods in Hypersonic Aerodynamics 473

for the Hermes configuration are shown in fig. 13. This is of particular importance
because it allows the definition of a generic form which contains the same physical
characteristics as the complex configuration. In the present case, e.g. the winglet,
shock interaction can be approximated by an oblique shock impinging on a swept
cylinder since the streamlines hitting the winglet do not originate from the nose region where vorticity is generated in the entropy layer.
8.3 2D Ramp
Some computational results for the 2D ramp, which was a test case for the Antibes
workshop are shown in figures 14 to 16. Figs. 14 and 15 show the iso-mach and iso-Cp
lines for the 111.4 test case. Fig. 16 shows the Cp distribution along the surface for 2
grid sizes: 200x65 and 200x130. It can be seen that the separation bubble grows when
refining the grid. Grid converged solutions are very time consuming but absolutely
mandatory. The ramp calculations serve as a simplified case to analyze the physical
phenomenon related to flap efficiency and surface heating.
8.4 Quasi ID Nozzle Flow in Thermo and Chemical Nonequilibrium
Figures 17 to 19 show results for a flow in thermo-chemical nonequilibrium using a
quasi one dimensional code. A shock at Mach 20 is simulated in an axisymmetric
channel by adjusting the pressure ratio and geometry [25]. The upstream density is
changed corresponding to a height of 70 and 20 km. Fig. 17 shows the mass fractions
along the nozzle and through the shock. The relatively low height of 20 km leads to
a very stiff system of equations, which goes to equilibrium immediately downstream
of the shock where the relaxation time is very small. Figs. 18 and 19 show the temperature distribution for the 20 km and for the 70 km case respectively. The thermal
nonequilibrium region is much larger for the low density case since the relaxation times
of the vibrational temperatures are of the same order as the flow times, resulting in a
much less stiff system which is much easier to solve [16, 36].
8.5 Double Ellipsoid
A six block grid was used for the real gas Euler computation on the double ellipsoide as shown in figure 20. The picture shows the crisp shock resolution, due to the
grid alignment with the shock. Here solution adaptation was used [28]. In fig. 21 a
Navier-Stokes solution with real gas in equilibrium for a mach number of 25 is shown.
The figure shows the Mach number distribution at an angle of incidence of 30 degrees.
8.6 Hermes
Here some results on the more complex HERMES configuration will be shown. Fig.
22 shows the surface temperature distribution for mach 21 and 30 degrees angle of
attack. Red color depicts high temperature and blue stands for low temperature 2
2See

Footnote 1.

474 Computational Methods in Hypersonic Aerodynamics


The temperatures on the surface are the inviscid temperatures obtained by the EULER computations. A synthesis of Euler computations performed by different groups
using real gas as well as perfect gas codes as well as a comparison with experiments at
Mach 6.4 and Mach 10 is shown in fig. 23. It is interesting to see that for the perfect
gas computations at Mach 10 and Mach 25 the same pitching moment coefficient is
obtained even though it is not obvious that the Mach number independence principle
should hold. However, at high mach numbers the real gas effects produce a pitch
up moment as was also found for the Orbiter. In general, good results are obtained
both for the lift coefficient and pitching moments. Not only do the 3D calculations
from various groups show very good agreements, but computations are also in good
agreement with data from the S4 wind tunnel. Fig. 24 shows a cross section at the
last plane to study the bow shock winglet interaction.
8.7 Orbiter Space Shuttle
Fig. 25 shows a Navier-Stokes solution for the Orbiter at Mach 10. The 4 block grid
of fig. 8 has been used. A close up of the nose region is shown in figure 26. A detailed
report of the "rebuilding of the orbiter" is given in [37].
9 CONCLUSIONS AND OUTLOOK
An overview of European present and possibly future hypersonics activities has been
presented. A derivation of the governing physical equations including thermo-chemical
nonequilibrium has been given. Grid generation has been treated in some detail, pointing out its crucial role to CFD applications. A brief introduction to the important
role of parallel computing in aerospace has been made. Numerous examples of hypersonic flow calculations for inviscid and viscous cases, for 2D and 3D shapes as well as
for real and perfect gas have been shown. The question of code validation was also
briefly addressed. Future activities will concentrate on the rebuilding of aerodynamic
effects and aerothermodynamic effects of the Space Shuttle. The influence of transition and turbulence models will also be investigated.
ACKNOWLEDGEMENT
The authors wish to express their gratitude to E. Slachmuylders, Head of Mechanical Engineering Department, ESTEC, who was instrumental in the creation of the
Aerothermodynamics Section and actively supported the section's activities over the
last three years.

Computational Methods in Hypersonic Aerodynamics 475

Figure 1: A multiblock Navier-Stokes airfoil grid is shown. The grid consists of 30


equally sized blocks, demonstrating that load balancing for a massively parallel system
can be achieved by domain decomposition.

Figure 2: A six block grid used for Euler simulations for the double ellipsoid.

476 Computational Methods in Hypersonic Aerodynamics

Figure 3: The Hermes monoblock grid with singularity at the nose is represented by
a spherical CS (coordinate system). The grid therefore has a pole at the nose and a
singularity line extends from there to the outer surface.

Figure 4: The monoblock Space Shuttle grid is of the same topology as the monoblock
Hermes grid (Fig. 3).

Computational Methods in Hypersonic Aerodynamics 477

Figure 5: The Hermes seven block grid is singularity free, and also leads to a reduction
of surface points without any detrimental effect on the accuracy of the solution.

Figure 6: The logical structure of the seven block Hermes grid is shown. The shaded
areas denote the Hermes surface.

478 Computational Methods in Hypersonic Aerodynamics

Figure 7: A Hermes cross section for a sophisticated multiblock grid is shown. The
left part shows the Euler grid, while the right part gives the corresponding N-S grid.

Figure 8: In this figure the four block Space Shuttle grid is shown. The topology is
that of Fig. 5. Removing the singularity line results in a much better convergence
rate of the N-S solver.

Computational Methods in Hypersonic Aerodynamics 479

-_.
--

Figure 9: Module Xivis (X stands for X-Windows, i for interactive and vis for visualization) of Grid* has been used to visualize the four block grid .

....... ( l -.

-)
(....
OooIPDot.O t2SW8>
..."',. Cf)

~>

...... III

'"-,....

"h.atfl

'

1f11 tBIXI>

=: r~~oJ__)
~

-.

CO . . . . . . . . O.~

<o.o...u.)

(-4

(-0.""".

1"'..- _, _ _II

(-() 1.-0 C8Il"OJo4)

'(.1. . . . . . 11 . . '
,
<-4 J 0 W!II)3'

(00'-'11.>

to

0 e.13II!II:t)

<-0 Ooet . O.
(-0

-.- ..

~U.O.

Figure 10: Module Xivis has been used to generate a close up of the grid forming the
pods and the vertical fin of the Space Shuttle.

480 Computational Methods in Hypersonic Aerodynamics

Figure 11: Solution adaptive grid for a cone. Solution parameters, pressure and Mach
number distributions, have been used to adapt the grid to better capture the shock
as well as to better resolve the BL.

Figure 12: In general the engineer is not interested in a set of numbers but expects
direct hints where to improve his design, that is locations where special physical effects
take place have to be visualized.

Computational Methods in Hypersonic Aerodynamics 481

Figure 13: In this picture vizualization serves to find out whether streamlines from
the nose region could come close to the winglets.
esteclypa

Antibes workshop 1991


testcase 111.4
grid 200 x 130
Machnumber contour lines

0.30

0.25

0.20

0.15

0.10

0.05

0.00
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Figure 14: 2D ramp flow (testcase IlI.4, Antibes workshop 1991). Presented are Mach
contour lines. A grid of dimension 200x130 was used.

482 Computational Methods in Hypersonic Aerodynamics

Antibes workshop 1991

estectypa

t est case 111.4


grid 200 x 130
cp contour lines

0.30

0.25

0.20

0.15

0.10

0.05

0.00
0.0

0.1

0.2

0.3

0.4

0.6

0.5

0.7

Figure 15: 2D ramp flow depicting pressure coefficient, Cp, contour lines.

0.250

0.200

Antibes 1991
testcase 111.4
cp along the vvall
grid 200x130
grid 200x6E

0.150

0.100
I

0.050

__

0.000 ~__-L-=======::::~::::~==~====~==~~~L-L__~____~__~____i 0.00


0.50
0.25

Figure 16: 2D ramp flow showing Cp values along the wall for two grid sizes.

Computational Methods in Hypersonic Aerodynamics 483


Mach = 20, Altitude = 70 km
O.BOO
_0_0_0

p;fp

_oo~~~~~-o-oo.o,o~

--------N2

0.600

0.400

02

~__- - -.......- O

_____----N

0.200

NO
0.02

0.03

0.04

0.05

0.06

0.07

O.OB

nozzle length [m]

Figure 17: Mass fractions for quasi ID nozzle flow. The shape of the nozzle has been
taken from Chiang and Hoffmann [36].
Chiang/Hoffmann Nozzle: Mach = 20, Altitude = 20 km
10 euler flow in thermal and chemical nonequilibrium
translational and vibrational temperatures
r'

10000

r"

Temp. [K]
----<:r- T

BOOO

-<>-

TN2

-6-

TNO

-0- TO~

i
_.-- .-.-...... -._- ... -_..
..... i....

6000

.1..

4000

.....~ .....

.... ~ .....

....-! ..-

I
"- i ._. -t- ....

".-. t

'~'-,'-':'-.---

....

....

..

I
i

j ....

IT i

..,'

.. j

r-'-'-'-'i-'T-'!
I

2000

, ....

i;
I

o ~. .~.-l.~~j;..~.~~..~.._~....~__~_~~:~_~_-i;;~~;_;;_=.___; _.L...._i_J ___._i. ___i - j

__

.. .. __

0.025

0.050

0.075

nozzle length [m]

Figure 18: Temperature distribution for the "Chiang and Hoffmann" nozzle at a
height of 20 km.

484 Computational Methods in Hypersonic Aerodynamics


Chiang/Hoffmann Nozzle: Mach = 20, Altitude = 70 km
10 euler flow in thermal and chemical nonequilibrium
translational and vibrational temperatures
Temp. (K)

1.S000E4

---?- T

--<>-

TN2

-0- T02
-4-

TNO- __

1.2000E4

8.0000E3 ~

4.0000E3

O.OOOOEO
0.020

0.040

O.OSO

0.080

nozzle length [m]

Figure 19: Temperature distribution for the "Chiang and Hoffmann" nozzle at a
height of 70 km.

Figure 20: This picture shows the temperature distribution for the double ellipsoid.

----_. __.-._-_._---

--

Computational Methods in Hypersonic Aerodynamics 485

Figure 21: The same double ellipsoid as in Fig. 20 was used. A Navier-Stokes solution
for a Mach number of 25 with real gas (equilibrium) has been performed at a constant
wall temperature of 1500 K.

Figure 22: Here a comparison of perfect gas (left half) and real gas (right half) for
the Hermes configuration at Mach 21 is shown.

486 Computational Methods in Hypersonic Aerodynamics


0.50
0.45

CL 0.40

0.030

equilibrium gas 89x33x73 l


equilibrium gas 45x17x37
DLRlBraunschweig
perfect gas
89x33x73
wind tunnel data R3Ch
wind tunnel data S4MA
BASE DASSAULT
.. equilibrium gas 74x72x33 (2D planar)
o equilibrium gas 62x72x36 (3D volume) ESTEC/YPA
x equilibrium gas 90x72x38 (3D volume)

0.35

0.30

0.25

.,"';'"= ... "'''~'('~

0.20

0.020

0.15
0.10

<>

0.05
0.00

L....JL......I---I.-J......J...--'--'--L-.!-.l--L......I---I.-J......J...--'--'--'-.!-.L..-L......I---l--L---l---'

10

15

20

0.01 0

25

00

Figure 23: The picture shows computed (using 3D codes) and measured values of
the lift coefficient, CI, (upper curves) and the pitching moment coefficient, Cm (lower
curves) for the Hermes 0-0 configuration.

Figure 24: The Mach number distribution for the Hermes space plane
obtained from Equilibrium Euler calculations.

IS

shown,

Computational Methods in Hypersonic Aerodynamics 487

Figure 25: A Navier-Stokes solution for the Shuttle Orbiter is shown, using the four
block grid of Fig. 8.

Figure 26: Close-up of the nose region of the Shuttle Orbiter from figure 25

488 Computational Methods in Hypersonic Aerodynamics


REFERENCES
1. Dannenhoffer, J.F. III, 1991: A Block-Structuring Technique for General Geometries, AIAA-91-0145, 14 pp.
2. Visich, M. et aI., 1991: Advanced Interactive Grid Generation Using Rambo-4G,
AIAA-91-0799, 5 pp.
3. Gentzsch, W., Hauser, J., 1988: Mesh Generation on Parallel Computers i S.
Sengupta et al. (Eds.), Numerical Grid Generation in Computational Fluid
Dynamics, Pineridge Press, pp. 113-124.
4. J. Hauser et al.: Parallel Computing in Aerospace Using Multiblock Grids, Part
I: Application to Grid Generation, submitted for publication (preprint available
from the authors) 22 pp.
5. J. Hauser et al., 1986: Boundary Fitted Conformed Co-ordinate Systems for
Selected Two Dimensional Fluid Flow Problems. Part I, II, J. Num. Meth.
Fluids, Vol 6, pp. 507-539.
6. P.R. Eiseman, 1990: Interactive Grid Generation with Control Points, Computing Systems in Eng., Vol 1, Nos 2-4, pp. 293-304.
7. J.F. Thompson, 1990: General Structured Grid Generation Systems, in: Applications of Mesh Generation to Complex 3-D configurations, AGARD-CP-464, 8
pp.
8. J.L. Steger, 1989: Technical Evaluation Report on the Fluid Dynamics Panel
Specialists' Meeting on Application of Mesh Generation to Complex 3-D configurations, AGARD-AR-268.
9. J. Hauser, A. Vinckier: Recent Developments in Grid Generation in: Applications of Mesh Generation to Complex 3-D Configurations, AGARD-CP-464, 15
pp.
10. G. Fox et al., 1988: Solving problems on Concurrent Processors, Vol. 1, Prentice
Hall, 592 pp.
11. Gustafson, J.L. et aI., 1988: Development of Parallel methods for a 1024Processor Hypercube, SIAM, J. of Scientific and Statistical Computing, Vol.
9, No.4, July, pp. 609-638.
12. Chapman, G.T., 1988: An Overview of Hypersonic Aerothermodynamics, Communications in Applied Numerical Methods, Vol. 4, pp. 319-325.
13. Lin, A.: Parallel Numerical Algorithms for Fluid Dynamics Simulation, AIAA900333.

---"---

Computational Methods in Hypersonic Aerodynamics 489


14. Hanel, D., Krause, E.: Finite Approximations in Fluid Mechanics II, DFG Priority Research Program, Results 1986-1988. Wiesbaden: Vieweg.
15. Candler, G.: The Computation of Weakly Ionized Hypersonic Flows in ThermoChemical Nonequilibrium, Phd thesis. Dept. of Aeronautics and Astronautics,
Stanford University, 1988.
16. Yu, S.-T., McBride, B.J., Hsieh, K.-C., Shuen, J.-S.: Numerical Simulation of
Hypersonic Inlet Flows with Equilibrium or Finite-Rate Chemistry, AIAA Paper
88-0273.
17. Millikan, R.C., White, D.R.: Systematics of Vibrational Relaxation, The Journal of Chemical Physics, Vol.39,pp.3209-3213,Dec 1963.
18. Blottner, F.: Chemical Reacting Viscous Flow Program for Multi-Component
Gas Mixtures, Report No. SC-RR-70-754, Sandia Laboratories,Dec 1971.
19. Wilke, C.R.: A Viscosity Equation for Gas Mixtures, The Journal of Chemical
Physics, VoLl8, pp.517-519, April 1950.
20. Jameson, A. and S. Yoon, 1987: Lower-Upper Implicit Schemes with Multiple
Grids for the Euler Equations; AIAA Journal, Vol.25, No.7, pp.929-935.
21. Hauser, J. et al., 1988: Numerical Solution of 2D and 3D Compressible Navier
Stokes Equations Using Time Dependent Boundary Fitted Coordinate Systems;
First International Short Course on Modern Computational Techniques in Fluid
Mechanics, University of Catalonia, Spain, 21 - 24 March, 50 pp.
22. Hauser, J., Simon, H.D., 1992: Aerodynamic Simulation on Massively Parallel
Systems; in Parallel Computational Fluid Dynamics '91, eds. W. Reintsch et
al., Elsevier-North Holland.
23. Koppenwallner, G., 1988: Aerthermodynamik - Ein Schluessel zu neuen Transportgeraeten des Luft und Raumfahrt, Z. Flugwiss. Weltraumforschung 12, pp.
6-18
24. Hauser, J., Vinckier, A., S. Zemsch, H.G. Paap 1989: Hypersonic Flow with
Non-equilibrium Effects, presented at Supercomputing in Fluid Flow, 3-5 October 1989, Lowell, Massachusetts, U.S.A., 38pp
25. Walpot, 1., 1991: Quasi One Dimensional Inviscid Nozzle Flow in Vibrational
and Chemical Non-Equilibrium, diploma thesis, T.U. Delft, Dept. of Aerospace
Eng. and ESA-ESTEC Aerothermodynamics Section YPA, The Netherlands,
51pp.
26. Ivanov, B., 1989: Fundamentals of Physics (chapters 5-6), Mir Publishers
Moscow, 455 pp.

490 Computational Methods in Hypersonic Aerodynamics


27. Hauser, J., Williams, R.D., 1991: Strategies for Parallelizing a Navier-Stokes
Code on the Intel Touchstone Machines, submitted to J. Num. Methods in
Fluids, 7 pp. (preprint available).
28. Wong, H., Haeuser, J., 1991: Equilibrium Solution of the Euler and NavierStokes Equations around a Double Ellipsoidal Shape with Mono- and MultiBlocks Including Real Gas Effects in Aerodynamics for Space Vehicles, ed. W.
Berry, pp. 453-458, ESA SP318
29. Hornung, H., Sturtevant, B., 1990: Challenges for High-Enthalpy Gasdynamics
Research During the 1990's, Caltech Graduate Aeronautical Lab., Report FM
90-1, 23 pp.
30. Anderson, A., 1989: Hypersonic and High-Temperature Gas Dynamics, McGrawHill
31. Park, C., 1989: Non-Equilibrium Hypersonic Aerothermodynamics, Wiley-Interscience, New York.
32. John, E.A., 1984: Gas Dynamics, Allyn and Bacon Inc.
33. Hauser, J, et al., 1992: Grid generation and Computational Aspects of MultiBlock Viscous Space Shuttle Calculations, preprint available from the authors.
34. Li, Chien-peng, 1985: Numerical Procedure for Three Dimensional Hypersonic
Viscous Flow over Aerobrake Configuration, NASA Lyndon B. Johnson Space
Center, Houston, Texas.
35. Hauser, J, Cassuli, V., 1992: A Fast and Positive Scheme for Solution of the
Non-Equilibrium Navier-Stokes Equations, preprint available from the authors.
36. Chiang, T.L., Hoffmann, K.A.: Determination of Computational Timestep for
Chemically reacting Flows, AIAA-89-1855.
37. Hauser, J, et al., 1992: Aerothermodynamics Computations for the Space Shuttle Orbiter, submitted to AIAA Thermophysics Conference, June 1992, Nashville,
U.S.A.

INDEX
Aerodynamic design optimization, 354
Atom/molecule mixture, 315
Avogadro numbers, 24
Axisymmetric hyperboloid, 282
Blunt body flow, 212
Boltzmann numbers, 24
Boundary Layer,
- chemical non-equilibrium, 13
- flat plate, 13
- laminar, 265
- thin layer approximation, 32
Chemical
- equilibrium, 164
- reaction, 6, 92, 298
- time scale, 301
Chemistry models, 3
Continuum flow, 9
Crossflow instability, 255
Defect approach, 276
Diffusion
- thermal, 17, 22
- phenomena, 26
Dissociation, 9
Double ellipsoid, flow over, 414
Dufour effect, 20
Equilibrium chemistry simulation, 396
Finite-rate chemistry, 163
Finite-volume method, 120, 179
Flat plate flow, 210, 245
Flow phenomena, 3, 30
Flow regimes, 103
Fluxes, 11, 46
Flux-difference splitting method, 44
Flux-vector splitting method, 44, 100
Free molecular flow, 9
Gas mixtures, 295, 303, 345
Gas phase chemistry, 297
Gortler instability, 248
High temperature flow, 21

Ionization, 10
Laminar-turbulent Transition, 233
Linear stability, 234
Multigrid computation, 203
MUSCL approach, 46
Nomenclature, 25, 115
Nopzle-afterbody, 341
Orbital vehicles, 7
Parabolized Navier-Stokes equations, 33
- conical method, 34
Perfect gas, 1
Physical aspects, 1
Plane hyperbola, 283
Point-implicit relaxation scheme, 115
Radiation, 94
Rarefied gas dynamics, 81
Reacting gas mixtures, 21
Real gas effects, 293
Real gas flows, 305
Receptivity, 239
Reentry trajectories, 9
Reentry vehicles, 29
Riemann Solvers, 43, 46
Runge-Kutta time-stepping scheme, 52
Sensitivity study, 333, 356
Shock, 3
Soret effect, 4
Steges-Warming type algorithm, 186
Thermal non-equilibrium, 176
Thermo-chemical non-equilibrium, 167
Thermodynamic models, 154, 157, 160
Transition, 240, 251
TVD scheme, 101
Van Dyke's theory, 267
Van Leer's splitting scheme, 66, 188
Vibrational excitation, 9
Note: The page references indicated are
only representative and are not meant to
be exhaustive.

Mechanics
FLUID MECHANICS AND ITS APPLICATIONS
Series Editor: R. Moreau
Aims and Scope of the Series
The purpose of this series is to focus on subjects in which fluid mechanics plays a fundamental
role. As well as the more traditional applications of aeronautics, hydraulics, heat and mass transfer
etc., books will be published dealing with topics which are currently in a state of rapid development, such as turbulence, suspensions and multiphase fluids, super and hypersonic flows and
numerical modelling techniques. It is a widely held view that it is the interdisciplinary subjects that
will receive intense scientific attention, bringing them to the forefront of technological advancement. Fluids have the ability to transport matter and its properties as well as transmit force,
therefore fluid mechanics is a subject that is particularly open to cross fertilisation with other
sciences and disciplines of engineering. The subject of fluid mechanics will be highly relevant in
domains such as chemical, metallurgical, biological and ecological engineering. This series is
particularly open to such new multidisciplinary domains.
1. M. Lesieur: Turbulence in Fluids. 2nd rev. ed., 1990
ISBN 0-7923-0645-7
2. O. Metais and M. Lesieur (eds.): Turbulence and Coherent Structures. 1991
ISBN 0-7923-0646-5
3. R. Moreau: Magnetohydrodynamics. 1990
ISBN 0-7923-0937-5
4. E. Coustols (ed.): Turbulence Control by Passive Means. 1990
ISBN 0-7923-1020-9
5. A. A. Borissov (ed.): Dynamic Structure of Detonation in Gaseous and Dispersed Media.
1991
ISBN 0-7923-1340-2
6. K.-S. Choi (ed.): Recent Developments in Turbulence Management. 1991
ISBN 0-7923-1477-8
ISBN 0-7923-1668-1
7. BHR Group (ed.): Pipeline Systems. 1992
8. BHR Group (ed.): Fluid Sealing. 1992
ISBN 0-7923-1669-X
9. T.K.S. Murthy (ed.): Computational Methods in Hypersonic Aerodynamics. 1992
ISBN 0-7923-1673-8

Kluwer Academic Publishers - Dordrecht / Boston / London

Mechanics
SOLID MECHANICS AND ITS APPLICATIONS
Series Editor: G.M.L. Gladwell
Aims and Scope of the Series
The fundamental questions arising in mechanics are: Why?, How?, and How much? The aim of this
series is to provide lucid accounts written by authoritative researchers giving vision and insight in
answering these questions on the subject of mechanics as it relates to solids. The scope of the series
covers the entire spectrum of solid mechanics. Thus it includes the foundation of mechanics;
variational formulations; computational mechanics; statics, kinematics and dynamics of rigid and
elastic bodies; vibrations of solids and structures; dynamical systems and chaos; the theories of
elasticity, plasticity and viscoelasticity; composite materials; rods, beams, shells and membranes;
structural control and stability; soils, rocks and geomechanics; fracture; tribology; experimental
mechanics; biomechanics and machine design.

0/ Structural Optimization. 2nd rev.ed.,


1990
ISBN 0-7923-0608-2
2. 1.1. Kalker: Three-Dimensional Elastic Bodies in Rolling Contact. 1990
ISBN 0-7923-0712-7
3. P. Karasudhi: Foundations o/Solid Mechanics. 1991
ISBN 0-7923-0772-0
4. N. Kikuchi: Computational Methods in Contact Mechanics. (forthcoming)
ISBN 0-7923-0773-9
5. Y.K. Cheung and A.Y.T. Leung: Finite Element Methods in Dynamics. (forthcoming)
ISBN 0-7923-1313-5
6. I.F. Doyle: Static and Dynamic Analysis 0/ Structures. With an Emphasis on Mechanics and
ISBN 0-7923-1124-8; Pb 0-7923-1208-2
Computer Matrix Methods. 1991
7. 0.0. Ochoa and I.N. Reddy: Finite Element Modelling 0/ Composite Structures.
(forthcoming)
ISBN 0-7923-1125-6
8. M.H. Aliabadi and D.P. Rooke: Numerical Fracture Mechanics.
ISBN 0-7923-1175-2
9. J. Angeles and C.S. Lopez-Cajun: Optimization o/Cam Mechanisms. 1991
ISBN 0-7923-1355-0
10. D.E. Grierson, A. Franchi and P. Riva: Progress in Structural Engineering. 1991
ISBN 0-7923-1396-8
11. R.T. Haftka and Z. Giirdal: Elements o/Structural Optimization. 3rd rev. and expo ed. 1992
ISBN 0-7923-1504-9; Pb 0-7923-1505-7
12. 1.R. Barber: Elasticity. 1992
ISBN 0-7923-1609-6
1. R.T. Haftka, Z. Giirdal and M.P. Kamat: Elements

Kluwer Academic Publishers - Dordrecht / Boston / London

Mechanics
From 1990, books on the subject of mechanics will be published under two series:

FLUID MECHANICS AND ITS APPLICATIONS


Series Editor: R.J. Moreau

SOLID MECHANICS AND ITS APPLICATIONS


Series Editor: G.M.L. Gladwell
Prior to 1990, the books listed below were published in the respective series indicated below.
MECHANICS: DYNAMICAL SYSTEMS
Editors: L. Meirovitch and G.lE. Oravas
1. E.H. Dowell: Aeroelasticity of Plates and Shells. 1975
ISBN 90-286-0404-9
2. D.G.B. Edelen: Lagrangian Mechanics of Nonconservative Nonholonomic Systems.
1977
ISBN 90-286-0077-9
3. J.L. Junkins: An Introduction to Optimal Estimation of Dynamical Systems. 1978
ISBN 90-286-0067-1
4. E.H. Dowell (ed.), H.C. Curtiss Jr., R.H. Scanlan and F. Sisto: A Modern Course in
Aeroelasticity.
Revised and enlarged edition see under Volume 11
5. L. Meirovitch: Computational Methods in Structural Dynamics. 1980
ISBN 90-286-0580-0
6. B. Skalmierski and A. Tylikowski: Stochastic Processes in Dynamics. Revised and
enlarged translation. 1982
ISBN 90-247-2686-7
7. P.e. MUller and W.O. Schiehlen: Linear Vibrations. A Theoretical Treatment of Multidegree-of-freedom Vibrating Systems. 1985
ISBN 90-247-2983-1
8. Gh. Buzdugan, E. Mihililescu and M. Rade: Vibration Measurement. 1986
ISBN 90-247-3111-9
9. G.M.L. Gladwell: Inverse Problems in Vibration. 1987
ISBN 90-247-3408-8
10. G.!. Schueller and M. Shinozuka: Stochastic Methods in Structural Dynamics. 1987
ISBN 90-247-3611-0
11. E.H. Dowell (ed.), H.C. Curtiss Jr., R.H. Scanlan and F. Sisto: A Modern Course in
Aeroelasticity. Second revised and enlarged edition (of Volume 4). 1989
ISBN Hb 0-7923-0062-9; Pb 0-7923-0185-4
12. W. Szemplinska-Stupnicka: The Behavior of Nonlinear Vibrating Systems. Volume I:
Fundamental Concepts and Methods: Applications to Single-Degree-of-Freedom
Systems. 1990
ISBN 0-7923-0368-7
13. W. Szempliriska-Stupnicka: The Behavior of Nonlinear Vibrating Systems. Volume II:
Advanced Concepts and Applications to Multi-Degree-of-Freedom Systems. 1990
ISBN 0-7923-0369-5
Set ISBN (Vols. 12-13) 0-7923-0370-9

MECHANICS OF STRUCTURAL SYSTEMS


Editors: J.S. Przemieniecki and G.lE. Oravas
1. L. Fryoa: Vibration of Solids and Structures under Moving Loads. 1970

2. K. Marguerre and K. Wolfel: Mechanics of Vibration. 1979

ISBN 90-01-32420-2
ISBN 90-286-0086-8

Mechanics
3. E.B. Magrab: Vibrations of Elastic Structural Members. 1979
ISBN 90-286-0207-0
4. RT. Haftka and M.P. Kamat: Elements of Structural Optimization. 1985
Revised and enlarged edition see under Solid Mechanics and Its Applications, Volume I

5. J.R Vinson and RL. Sierakowski: The Behavior of Structures Composed of Composite
ISBN Hb 90-247-3125-9; Pb 90-247-3578-5
Materials. 1986
6. B.E. Gatewood: Virtual Principles in Aircraft Structures. Volume 1: Analysis. 1989
ISBN 90-247-3754-0
7. B.E. Gatewood: Virtual Principles in Aircraft Structures. Volume 2: Design, Plates,
Finite Elements. 1989
ISBN 90-247-3755-9
Set (Gatewood 1 + 2) ISBN 90-247-3753-2
MECHANICS OF ELASTIC AND INELASTIC SOLIDS
Editors: S. Nemat-Nasser and G.lE. Oravas
1. G.M.L. Gladwell: Contact Problems in the Classical Theory of Elasticity. 1980
ISBN Hb 90-286-0440-5; Pb 90-286-0760-9
2. G. Wempner: Mechanics of Solids with Applications to Thin Bodies. 1981
ISBN 90-286-0880-X
3. T. Mura: Micromechanics of Defects in Solids. 2nd revised edition, 1987
ISBN 90-247-3343-X
4. R.G. Payton: Elastic Wave Propagation in Transversely Isotropic Media. 1983
ISBN 90-247-2843-6
5. S. Nemat-Nasser, H. Abe and S. Hirakawa (eds.): Hydraulic Fracturing and Geothermal Energy. 1983
ISBN 90-247-2855-X
6. S. Nemat-Nasser, RJ. Asaro and G.A. Hegemier (eds.): Theoretical Foundation for
Large-scale Computations of Nonlinear Material Behavior. 1984 ISBN 90-247-3092-9
7. N. Cristescu: Rock Rheology. 1988
ISBN 90-247-3660-9
8. G.I.N. Rozvany: Structural Design via Optimality Criteria. The Prager Approach to
Structural Optimization. 1989
ISBN 90-247-3613-7
MECHANICS OF SURFACE STRUCTURES
Editors: W.A. Nash and G.lE. Oravas
1. P. Seide: Small Elastic Deformations of Thin Shells. 1975
ISBN 90-286-0064-7
2. V. Panc: Theories of Elastic Plates. 1975
ISBN 90-286-0 1O4-X
3. J.L. Nowinski: Theory ofThermoelasticity with Applications. 1978
ISBN 90-286-0457-X
4. S. Lukasiewicz: Local Loads in Plates and Shells. 1979
ISBN 90-286-0047-7
5. C. Fin: Statics, Formfinding and Dynamics of Air-supported Membrane Structures.
1983
ISBN 90-247-2672-7
6. Y. Kai-yuan (ed.): Progress in Applied Mechanics. The Chien Wei-zang Anniversary
ISBN 90-247-3249-2
Volume. 1987
7. R. Negruliu: Elastic Analysis of Slab Structures. 1987
ISBN 90-247-3367-7
8. J.R. Vinson: The Behavior of Thin Walled Structures. Beams, Plates, and Shells. 1988
ISBN Hb 90-247-3663-3; Pb 90-247-3664-1

Mechanics
MECHANICS OF FLUIDS AND TRANSPORT PROCESSES
Editors: R.J. Moreau and G.JE. Oravas
1. J. Happel and H. Brenner: Low Reynolds Number Hydrodynamics. With Special
Applications to Particular Media. 1983
ISBN Hb 90-01-37115-9; Pb 90-247-2877-0
2. S. Zahorski: Mechanics of Viscoelastic Fluids. 1982
ISBN 90-247-2687-5
3. J.A. Sparenberg: Elements of Hydrodynamics Propulsion. 1984 ISBN 90-247-2871-1
ISBN 90-247-2999-8
4. B.K. Shivamoggi: Theoretical Fluid Dynamics. 1984
5. R. Timman, A.J. Hermans and G.C. Hsiao: Water Waves and Ship Hydrodynamics. An
Introduction. 1985
ISBN 90-247-3218-2
6. M. Lesieur: Turbulence in Fluids. Stochastic and Numerical Modelling. 1987
ISBN 90-247-3470-3
7. L.A. Lliboutry: Very Slow Flows of Solids. Basics of Modeling in Geodynamics and
Glaciology. 1987
ISBN 90-247-3482-7
8. B.K. Shivamoggi: Introduction to Nonlinear Fluid-Plasma Waves. 1988
ISBN 90-247-3662-5
9. V. Bojarevics, Ya. Freibergs, E.I. Shilova and E.V. Shcherbinin: Electrically Induced
Vortical Flows. 1989
ISBN 90-247-3712-5
10. J. Lielpeteris and R. Moreau (eds.): Liquid Metal Magnetohydrodynamics. 1989
ISBN 0-7923-0344-X
MECHANICS OF ELASTIC STABILITY
Editors: H. Leipholz and G.JE. Oravas
1. H. Leipholz: Theory of Elasticity. 1974
ISBN 90-286-0193-7
2. L. Librescu: Elastostatics and Kinetics of Aniosotropic and Heterogeneous Shell-type
ISBN 90-286-0035-3
Structures. 1975
3. C.L. Dym: Stability Theory and Its Applications to Structural Mechanics. 1974
ISBN 90-286-0094-9
4. K. Huseyin: Nonlinear Theory of Elastic Stability. 1975
ISBN 90-286-0344-1
5. H. Leipholz: Direct Variational Methods and Eigenvalue Problems in Engineering.
1977
ISBN 90-286-0106-6
6. K. Huseyin: Vibrations and Stability of Multiple Parameter Systems. 1978
ISBN 90-286-0136-8
7. H. Leipholz: Stability of Elastic Systems. 1980
ISBN 90-286-0050-7
8. V.V. Bolotin: Random Vibrations of Elastic Systems. 1984
ISBN 90-247-2981-5
9. D. Bushnell: Computerized Buckling Analysis of Shells. 1985
ISBN 90-247-3099-6
10. L.M. Kachanov: Introduction to Continuum Damage Mechanics. 1986
ISBN 90-247-3319-7
11. H.H.E. Leipholz and M. Abdel-Rohman: Control of Structures. 1986
ISBN 90-247-3321-9
12. H.E. Lindberg and A.L. Florence: Dynamic Pulse Buckling. Theory and Experiment.
1987
ISBN 90-247-3566-1
13. A. Gajewski and M. Zyczkowski: Optimal Structural Design under Stability Constraints.1988
ISBN 90-247-3612-9

Mechanics
MECHANICS: ANALYSIS
Editors: V.J. Mizel and G.tE. Oravas
1. M.A. Krasnoselskii, P.P. Zabreiko, E.I. Pustylnik and P.E. Sbolevskii: Integral
ISBN 90-286-0294-1
Operators in Spaces of Summable Functions. 1976
2. V.V. Ivanov: The Theory of Approximate Methods and Their Application to the
Numerical Solution of Singular Integral Equations. 1976
ISBN 90-286-0036-1
3. A. Kufner, O. John and S. Pucik: Function Spaces. 1977
ISBN 90-286-0015-9
4. S.G. Mikhlin: Approximation on a Rectangular Grid. With Application to Finite
Element Methods and Other Problems. 1979
ISBN 90-286-0008-6
5. D.G.B. Edelen: Isovector Methods for Equations of Balance. With Programs for
Computer Assistance in Operator Calculations and an Exposition of Practical Topics of
the Exterior Calculus. 1980
ISBN 90-286-0420-0
6. R.S. Anderssen, F.R. de Hoog and M.A. Lukas (eds.): The Application and Numerical
Solution of Integral Equations. 1980
ISBN 90-286-0450-2
7. R.Z. Has'minskiI: Stochastic Stability of Differential Equations. 1980
ISBN 90-286-0100-7
8. A.I. Vol'pert and S.I. Hudjaev: Analysis in Classes of Discontinuous Functions and
Equations of Mathematical Physics. 1985
ISBN 90-247-3109-7
9. A. Georgescu: Hydrodynamic Stability Theory. 1985
ISBN 90-247-3120-8
10. W. Noll: Finite-dimensional Spaces. Algebra, Geometry and Analysis. Volume I. 1987
ISBN Hb 90-247-3581-5; Pb 90-247-3582-3
MECHANICS: COMPUTATIONAL MECHANICS
Editors: M. Stern and G.tE. Oravas
1. T.A. Cruse: Boundary Element Analysis in Computational Fracture Mechanics. 1988

ISBN 90-247-3614-5
MECHANICS: GENESIS AND METHOD
Editor: G.lE. Oravas
1. P.-M.-M. Duhem: The Evolution of Mechanics. 1980

ISBN 90-286-0688-2

MECHANICS OF CONTINUA
Editors: W.O. Williams and G.tE. Oravas
1. C.-C. Wang and C. Truesdell: Introduction to Rational Elasticity. 1973
ISBN 90-01-93710-1
2. P.J. Chen: Selected Topics in Wave Propagation. 1976
ISBN 90-286-0515-0
3. P. Villaggio: Qualitative Methods in Elasticity. 1977
ISBN 90-286-0007-8

Mechanics
MECHANICS OF FRACTURE
Editors: G.c. Sih
1. G.C. Sih ted.): Methods of Analysis and Solutions of Crack Problems. 1973
ISBN 90-01-79860-8
2. M.K. Kassir and G.c. Sih (eds.): Three-dimensional Crack Problems. A New Solution
of Crack Solutions in Three-dimensional Elasticity. 1975
ISBN 90-286-0414-6
3. G.C. Sih (ed.): Plates and Shells with Cracks. 1977
ISBN 90-286-0146-5
4. G.C. Sih (ed.): Elastodynamic Crack Problems. 1977
ISBN 90-286-0156-2
5. G.C. Sih (ed.): Stress Analysis of Notch Problems. Stress Solutions to a Variety of
Notch Geometries used in Engineering Design. 1978
ISBN 90-286-0166-X
6. G.C. Sih and E.P. Chen (eds.): Cracks in Composite Materials. A Compilation of Stress
ISBN 90-247-2559-3
Solutions for Composite System with Cracks. 1981
7. G.c. Sih (ed.): Experimental Evaluation of Stress Concentration and Intensity Factors.
Useful Methods and Solutions to Experimentalists in Fracture Mechanics. 1981
ISBN 90-247-2558-5
MECHANICS OF PLASTIC SOLIDS
Editors: J. Schroeder and G.lE. Oravas
1. A. Sawczuk (ed.): Foundations of Plasticity. 1973
ISBN 90-01-77570-5
2. A. Sawczuk (ed.): Problems of Plasticity. 1974
ISBN 90-286-0233-X
3. W. Szczepiflski: Introduction to the Mechanics of Plastic Forming of Metals. 1979
ISBN 90-286-0126-0
4. D.A. Gokhfeld and O.F. Chemiavsky: Limit Analysis of Structures at Thermal Cycling.
1980
ISBN 90-286-0455-3
5. N. Cristescu and I. Suliciu: Viscoplasticity. 1982
ISBN 90-247-2777-4

Kluwer Academic Publishers - Dordrecht / Boston / London

Potrebbero piacerti anche