Dynamics of Compressive Fluid Flow PDF
Dynamics of Compressive Fluid Flow PDF
Dynamics of Compressive Fluid Flow PDF
and Thermodynamics of
COMPRESSIBLE FLUID
FLOW
by
ASCHER H. SHAPIRO
Volume I
[email protected] : & 9 F f v-
The Dynamics
and Thermodynamics of
COMPRESSIBLE FLUID
FLOW
BY
ASCHER H . SHAPIRO
Professor of Mechanical Engiiredng
Massachusetts Institute of Technology
w o VOLUMES
VOLUME I
I S B N 0 471 06691-5
10 9
TO THE MEMORY OF MY FATHEH.
BERNARD SHAPIRO
PREFACE
During the past two decades a rapid growth of interest in the motion
of compressible fluids has accompanied developments in high-speed
flight, jet engines, rockets, ballistics, combustion, gas turbines, ram jets
and other novel propulsive mechanisms, heat transfer a t high speeds,
and blast-wave phenomena. My purpose in writing this book is to make
available to students, engineers, and applied physicists a work on com-
pressible fluid motion which would be suitable as an introductory text
in the subject as well as a reference work for some of its more advanced
phases. The choice of subject matter has not been dictated by any
particular field of engineering, but rather includes topics of interest to
aeronautical engineers, mechanical engineers, chemical engineers, ap-
plied mechanicians, and applied physicists.
I n selecting material from the vast literature of the field the basic
objective has been to make the book of practical value for engineering
purposes. To achieve this aim, I have foilowed the philosophy that the
most practical approach to the subject of compressible fluid mechanics
is one which combines theoretical analysis, clear physical reasoning, and
empirical results, each leaning on the other for mutual support and ad-
vancement, and the whole being greater than the sum of the parts.
The analytical developments of this book comprise two types of treat-
ments: those leading to design methods and those leading to exemplary
methods. The design methods are direct and rapid, and easily applied
to a variety of problems. Therefore, they are suited for use in the engi-
neering office. The discussions of these design methods are detailed and
illustrative examples are often given. The exemplary methods, on the
other hand, comprise those theoretical analyses which are time consum-
ing, which generally require mathematical invention, and which are not
easily applied to a variety of problems. Such methods are primarily of
value for ~ieldingdetailed answers to a small number of typical prob-
lems. Although they are not in themselves suitable for the engineering
office, the examples which they permit to be worked out often provide
important information about the behavior of fluids in typical situations.
Thus they serve as guides to the designer in solving the many complex
problems where even the so-called design met,hods are not sufficient.
The treatment of exemplary methods in this book usually consists of a
brief outline of the method, together with a presentation of those results
obtained by the method which illuminate significant questions concern-
V
vi PREFACE
ing fluid motion and which help to form the vital "feel" so desired by
designers.
In keeping with the spirit of the several foregoing remarks, all
the important results of the book have been reduced to the form
of convenient charts and tables. Unless otherwise specified, the
charts and tables are for a perfect gas with a ratio of specific heats (k)
of 1.4.
In those parts of the book dealing with fundamentals, emphasis is
placed on the introduction of new concepts in an unambiguous manner,
on securing a clear physical understanding before the undertaking of an
analysis, on the rigorous application of physical laws, and on showing
fruitful avenues of approach in analytical thinking. The remaining part
of the work proceeds a t a more rapid pace befitting the technical ma-
turity of advanced students and professionals.
The work is organized in eight parts. Part I sets forth the basic con-
cepts and principles of fluid dynamics and thermodynamics from which
the remainder of the book proceeds and also introduces some funda-
mental concepts peculiar to compressible flows. I n Part I1 is a discus-
sion of problems accessible by the most simple picture of fluid motion
-the one-dimensional analysis. Part I11 constitutes a summary of the
basic ideas and concepts necessary for the succeeding chapters on two-
and three-dimensional flow. Parts IV, V, and VI then present in order
comprehensive surveys of subsonic flows, of supersonic flows (including
hypersonic flow), and of mixed subsonic-supersonic flows. In Part VII
is an exposition of unsteady one-dimensional flows. Part VlII is an
examination of the viscous and heat conduction effects in laminar and
turbulent boundary layers, and of the interaction between shock waves
and boundary layers. For those readers not already familiar with it,
the mathematical theory of characteristic curves is briefly developed in
Appendix A. Appendix B is a collection of tables which facilitate the
numerical solution of problems.
The "References and Selected Bibliography" a t the end of each c h a p
ter will, it is hoped, be a helpful guide for further study of the volumi-
nous subject. Apart from specific references cited in each chapter, the
lists include general references appropriate to the subject matter of each
chapter. The choice of references has been based primarily on clarity,
on completeness, and on the desirability of an English text, rather than
on historical priority.
My first acknowledgment is to Professor Joseph H. Keenan, to whom
1 owe my first interest in the subject, and who, as teacher, friend, and
colleague, has been a source of inspiration and encouragement.
In an intangible yet real way I am indebted to my students, who have
made teaching a satisfying experience, and to my friends and colleagues
PREFACE tril
at the Massachusetts Institute of Technology who contributed the cli-
mate of constructive criticism so conducive to creative effort.
Many individuals and organizations have been cooperative in supply-
ing me with helpful material and I hope that I have not failed to ac-
knowledge any of these a t the appropriate place in the text. The Na-
tional Advisory Committee for Aeronautics and the M.I.T. Gas Turbine
Laboratory have been especially helpful along these lines.
I was fortunate in being able to place responsibility for the important
work of the drawings in the competent hands of Mr. Percy H. Lund,
who, with Miss Prudence Santoro, has been most cooperative in this
regard.
For help with the final revision and checking of the manuscript I wish
to give thanks to Dr. Bruce D. Gavril and Dr. Ralph A. Burton.
Finally, but by no means least, I must express a word of appreciation
to Sylvia, and to young Peter, Mardi, and Bunny, who, one and all,
made it possible for me to escape from the office into the somewhat less
trying atmosphere of the home, and there to carry this work forward
to its completion.
ASCRERH. SHAPIRO
Arlington, Mass.
April 4, 1953
CONTENTS
VOLUME I
Part I . Background
CHAPTER PAGE
1 FOUNDATIONS OF FLUID DYNAMICS 3
Properties of the Continuum. Systems and Control Volumes.
Conservation of Mass. Momentum Theorem. Theorem of Mo-
ment of Momentum. Units and Dimensions.
2 FOUNDATIONS OF THERMODYNAMICS 23
The First Law of Thermodynamics. The Second Law of Thermo-
dynamics. Thermodynamic Properties of the Continuum. The
First Law for a Control Volume., The Second Law fot a Control
Volume. The Perfect Gas.
11 HODOGRAPH
METHOD FOR TWO-DIMENSIONAL, SUBSONIC
FLOW 336
Derivation of the Hodograph Equations. The Tangent-Gas Ap-
proximation. The Karman-Tsien Pressure Correction Formula.
Calculation of Profile Shape Correction. Extension of Karman-
Tsien Method. Miscellaneous Examples.
CONTENTS
CHAIT E R
12 MISCELLANEOUS METHODS
A N D RESULTS FOR TPio-DIMEN-
SIONAL, SUBSONIC FLOW
The Rayleigh-Janzen Method of Expansion in Series of the Mach
Number. The Prandtl-Glauert Method of Expansion in Series of
a Shape Parameter. Relaxation Method. Some Measured Effects
of Compressibility in Subsonic Flow. The Streamline Curvature
Method.
THREE-DIMENSIONAL, SUBSONIC FLOW
Gothert's Rule for Uniform Flow with Small Perturbations. Flow
Past Ellipsoids. Bodies of Revolution. Spheres. Wings of Finite
Span. Sweptback Wings. Sweptback Wings of Finite Span.
Appendix
THEORY OF CHARACTERISTICS
The Characteristic Curves. Method of Constructing Character-
istic Curves. Simple Waves.
VOLUME II
Part V . Supersonic Flow (Continued)
CHAPTER
17 AXIALLY SYMMETRIC SUPERSONIC FLOW
Exact Solution for Flow Past a Cone. Linear Theory for Slender
Bodies of Revolution. Method of Characteristics. Miscellaneous
Experimental Results.
19 HYPERSONIC FLOW
Similarity Laws for Hypersonic Flow. Oblique Shock Relations
for Hypersonic Flow. Simple-Wave Expansion Relations for Hy-
personic Flow. Hypersonic Performance of Two-Dimensional
Profiles. Hypersonic Performance of Bodies of Revolution. Ex-
perimental Results.
21 TRANSONIC FLOW
The Transonic Similarity Law. Applications of the Transonic
Similarity Law. Flow in Throat of Converging-Diverging Nozzle.
Relaxation Method. Transonic Flow Past a Wavy Wall. Flow
a t Mach Number Unity. Slopes of Force Coefficients a t M, = 1.
Transonic Flow Past Wedge Nose.
NOMENCLATURE
acceleration u component of velocity in x-
area direction
area vector ZJ component of velocity in y-
force direction
force vector '0 volume
magnitude of body force per V speed
unit mass V velocity
constant of proportionality in w mass rate of flow
Newton's second law x, y, z Cartesian coordinates
mass
moment of a force, or torque
normal force per unit area, or Y angle
pressure '/ coefficient of viscosity
magnitude of radius vector P mass density at a point
radius vector T tangential force per unit area,
time or shear stress
(4 (b)
FIG.1.1. Definition of density at a point.
useful concept.
Fluid Velocity at a Point. The fluid velocity a t a point is quite inde-
pendent of the instantaneous velocity of the molecule nearest that point.
Rather we consider the motion of the center of gravity of the volume
62,' (Fig. l.lb) instantaneously surrounding that point, and define the
jluid velocity at the point P as the instantaneous velocity of this center of
gravity. Thus the fluid velocity a t a point is the instantaneous velocity
of the fluid particle which a t that moment is passing through the point.
By jluid particle we mean here a small mass of fluid of fixed identity and
of size comparable with 62,'.
Whereas density a t a point is a scalar quantity, fluid velocity a t
a point is a vector. After the introduction of a coordinate system,
it is therefore possible to resolve the vector velocity into three scalar
components.
Art. 1.2 PROPERTIES OF THE CONTINUUM 7
STREAMLINES. At any instant, every point of a fluid continuum has
a corresponding fluid velocity vector. The curves which are everywhere
tangent to the velocity vector are known as the instantaneous streamlines
of the flow, and comprise one of the most common and most useful graph-
ical representations of the flow. When we speak of the flow pattern, we
often mean the streamline picture.
In unsteady flows the streamline pattern changes from instant to
instant. When the flow is steady, however, the streamlines are constant in
time and represent also the path lines, or the trajectories of fluid particles.
STREAM TUBE. Consider an elementary area dA normal to the veloc-
ity vector a t some point. The streamlines passing through the circum-
ference of d A form a surface which is called a stream tube. Since there
is no velocity component normal to the streamline, the walls of a stream
tube may be thought of as being impervious to flow.
Stress at a Point. Consider a plane passing in a given direction
through point P of the continuous medium of Fig. 1.2, and imagine a
I Domain of I
(4 (b)
FIG.1.2. Definition of normal stress at a point.
Both the pressure and the shear stress are, of course, vector quantities.
The area 6A' and volume 6V1, i t should be noted, have comparable
dimensions.
HYDROSTATIC PRESSURE. We now raise the question whether a t the
point P the pressure, or normal stress, depends upon the orientation of
the plane 6A.
To answer this question, let us first consider an inviscid fluid, that is,
a fluid in which no viscous stresses (tangential or shear stresses) exist
(1 -
even though there is relative motion within the fluid. Suppose for
I
I
I
I
I
I
\
_.
,------
Connituum-l
&
\ pxdy
/
dy
--- \
- 7--J
pydx
\
\
\'
\I
I
I
mathematical simplicity that only
two dimensions are considered, and
let us then investigate the dynamics
of an infinitesimally small triangu-
lar piece of fluid of unit depth sur-
rounding P in Fig. 1.3. Then, since
shear stresses are assumed absent,
only normal stresses, or pressures,
act on the three faces. The forces
'.
.----
\
_---__----/' x
FIG.1.3. Normal stress forces acting
exerted by the fluid outside the
triangle on the fluid within the
at a point. triangle are shown in the sketch,
where p, denotes the pressure a t
point P in the x-direction, p, denotes the pressure a t point P in the
y-direction, and p, denotes the pressure a t point P in the 7-direction.
Let g, and g, be the components of body force (e.g., gravity force)
acting on the fluid per unit mass. Now we may write Newton's second
law in the x-direction, as follows:
Relations like Eqs. 1.6 and 1.7 may be written for all continuum
properties thus far defined.
STEADY FLOW. Many engineering problems refer to a steady condi-
tion of operation. Specifically we define steady$ow as a condition where
at, each point of space there is no variation of any property with respect
to time. Using the density for illustration, we may then write
(b)
Fro. 1.5. Flow through a control volume.
+
I and 111, and a t time t dt it occupies spaces I and 11. Thus, since
the mass of the system is conserved, we write
where ml, means the mass of fluid in space I a t time t, and so on. A
simple rearrangement then gives
The first term represents the time rate of change of mass within space
I. But, as dt goes to zero, space I coincides with the control volume, and
so, in the limit,
where m,.,, denotes the instantaneous mass within the control volume.
The third term may be written
S
denoted for convenience by dwout. The integral sign in dwoutsignifies
that the mass rate of flow is to be summed up for all elements of control
surface area dAout through which fluid leaves the control volume.
Similar reasoning yields
which states that the rate of accumulation of mass within the control
volume is equal to the excess of the incoming rate of flow over the out-
going rate of flow.
Continuity Equation. For detailed computations, we note that at any
instant
With the help of Fig. 1.5b, we may express the mass rate of flow in
where the last integral is a surface integral summed up over the entire
control surface.
Steady Flow. When the flow is steady, the identity of the fluid within
the control volume changes continuously, but the total mass remains
constant. Or, mathematically, ap,Qt is zero for each element of control
volume. Then Eq. 1.9 states that the incoming and outgoing mass
rates of flow are identical:
I PZVZAZ (1-12)
PIVIA=
where A~ represents the cross-sec- FIG.1.6. Onedimensional flow
16 FOUNDATIONS OF FLUID DYNAMICS Ch. 1
rial occupying the control volume a t time t. For the sake of generality
there is also shown protruding into the control volume some sort of
obstacle, such as a strut, screen, flameholder, guide vane, or turbo-
machine blade. For the control volume chosen, that part of the obstacle
lying within the control volume is part of the system whose motion dur-
ing the time dt is being studied.
Newton's second law is a vector relation. Considering the x-direction,
we write, for the system
where the left-hand side represents the algebraic sum of the x-forces
Art. 1.5 MOMENTUM THEOREM 17
acting on the system during the time interval dt, and the right-hand side
represents the time rate of change of the total x-momentum of the
system. The right-hand side may, however, be expressed as
and note that a similar expression applies to the incoming flow. The
term
S V, dw which appears here is called the flux of x-momentum.
Now, with the expressions developed above, the dynamic relation
may be written
where p and V, are the local density and x-velocity, respectively, of the
fluid element of volume d o within the control volume. Then, using the
expression for dw developed in Art. 1.4, the momentum theorem is
written
ZF, = J
c.v.
'
at
bvz)do +Jpvnvz dA0.t -JpvnvZ dAin (1-14)
18 FOUNDATIONS OF FLUID DYNAMICS Ch. 1
Two similar relations apply to the y- and z-directions. By using vector
notation, the momentum theorem for a control volume may be repre-
sented by a single equation,
a
2F =-
at J,..PV dl) +$ dV.dA)V
C.8.
(1.15)
form is not valid since there must be inserted additional terms to take
account of the centrifugal and Coriolis accelerations. Since the mo-
mentum theorem derives from the law of motion, a similar conclusion
applies, namely that in Eq. 1.15 the velocities must be measured with
respect to the earth, or with reference to a coordinate system moving
a t constant velocity with respect to the earth.
I(&
where r is the radius vector from an arbitrary origin of moments. In
Cartesian notation, and referring
Y
to Fig. 1.8, this is written
d
M, = ZrFt =- (ZmrVJ
at
where r is the radius in the xy-plane /
which states that the algebraic sum of the moments is equal to the time
rate of change of moment of momentum within the control volume plus
the excess of outgoing flux of moment of momentum over the correspond-
ing incoming flux.
The remarks in Art. 1.5 concerning forces are evidently applicable
also to the moments of the forces.
-
could be measured were totally unrelated, and so Newton could state
his law only as a proportionality, F ma, or
-
Mass Length Time
- Force
-
Ibm ft sec lbf 32.174 lbm f t2
lbf sec
Ibm ft
Ibm ft sec poundal 1
poundal sec2
I slug ft
slug ft sec Ibf
lbf sec2
The last system, which is the one employed in this book, is one where
go has the numerical magnitude unity, and thus one for which unit force
gives unit mass unit acceleration. For such a system of units, we may
eliminate go from the mathematical statement of the law, which is
tantamount to eliminating one of the primary units of measure and
defining i t instead in terms of the remaining three. Thus, writing as
we do,
F = m
Art. 1.7 PROBLEMS
Since we usually employ the slug a s mass unit for numerical calcula-
tions, but often refer t o results in terms of pound masses, it is convenient
t o remember t h a t
1 slug = 32.174 ibm
PROBLEMS
1.1. Consider the frictionless, steady flow of a compressible fluid in an in-
finitesimal stream tube.
(a) Demonstrate by the continuity and momentum theorems that
- + - + dA
dp
- = o dV
P A V
dp+pVdV+pgdz= 0
(b) Determine the integrated forms of these equations for an incompressible
fluid.
v AtdA
P Gmvity 2 zrdz
P
WV
Nozzle -
1.4. The sketch shows a vane with a turning angle P which moves with a
steady speed U. The vane receives a jet which leaves a fixed nozzle with speed V.
(a) Assuming that the vane is mounted on rails as shown in the sketch, show
that the work done against the restraining force is a maximum when U/V = M.
(b) Assuming that there are a large number of such vanes attached to a rotat-
ing wheel moving with peripheral speed U, show that the work delivered to the
wheel is a maximum when U/V = x.
1.6. In an experiment to determine drag, a circular cylinder of diameter d
--
=-- "0
----- "0
C
---+ --
"0
-
was immersed in a steady, twodimen-
sional, incompressible flow. Measure-
ments of velocity and pressure were
a made at the boundaries of the control
avo B i d 4d
surface shown. The pressure was found
to be uniform over the entire control
t;
--
surface. The x-component of velocity a t
I ,
."I -- ----+ --
Control Surface
Vo "a
1
Vo
-- the control surface boundary was
approximately as indicated by the
sketch.
PROB. 1.5. From the measured data, calculate
the drag coefficient of the cylinder,
based on the projected area and on the free stream dynamic head, y$Vo2.
Drag Force per Unit Length
CD =
$pV&
Chapter 2
FOUNDATIONS OF THERMODYNAMICS
NOMENCLATURE
area 4' vector heat flux per unit area
specific heat a t constant pressure 2 rate of heat transfer
specific heat a t constant volume R gas constant
internal energy per unit mass 6i universal gas constant
internal energy of a system in s entropy per unit mass
general S entropy
local acceleration of gravity T absolute temperature
enthalpy per unit mass u same as U, except for a unit
ratio of specific heats mass
mass U that part of the internal energy
pressure of a system independent of
power motion, gravity, electricity,
heat capillarity, and magnetism
heat transfer per unit time per v specific volume
unit area in the x-direction 2, volume
23
24 FOUNDATIONS OF TITERMODYNAMICS Ch.2
NOMENCLATURE--Continued
V speed thermal effieicncy
w mass rate of flow X coefficient of thermal conduc-
W work; molecular weight tivity
z elevation above given datum p der~sity
which states that the algebraic net heat received by the system during a
Art. 2.3 rHE SECOND LAW OF THERMODYNAMICS 29
cycle is equal to the algebraic net work done by the system during the
cycle.
Adiabatic Process. It is a matter of experience that when barriers,
called heat insulators, are placed between a system and surroundings
a t different temperatures, the heat Q reckoned from Eq. 2.3 becomes
very small, and it is not difficult to extrapolate in one's mind to the
condition where the insulation is so effective as to make the heat zero.
Such a process is said to be adiabatic. Since Q is also zero in an equal-
temperature process, it follows that an equal-temperature process ir
also an adiabatic process.
Metal
(a) (b)
FIG. 2.1. Conduction of heat from low temperature to high temperature permits
perpetual motion of second kind.
Thus we have proved that our oft perceived observation of the irre-
versibility of heat conduction from a high-temperature to a low-tem-
perature body is in fact a logical consequence of the Second Law. The
proposition stated above may be similarly proved for irreversible
processes in general, and this suggests an alternative definition of
irreversibility :
A process is irreversible i f the undoing of all the efects of the process
would make possible the construction of a perpetual motion machine of the
second kind.
Corollaries of the Second Law. The following corollaries of the Second
Law, comprising the most important results for its practical utilization,
stem by logical processes from its statement. Again we follow closely
the development of Keenan.(')
COROLLARY 1. NO heat engine operating between two heat reservoirs
of fixed and uniform temperature can have a greater thermal efficiency
than a reversible heat engine which operates between the same two
reservoirs, for it can be shown that otherwise perpetual motion of the
second kind would be possible.
32 FOUNDATIONS OF THERMODYNAMICS Ch. 2
of the system during an elementary part of the cycle, and T is the corre-
sponding absolute temperature a t that part of the boundary. The
integral is to be summed for all parts of the boundary and over the
entire cycle.
From this it follows that in going from a given state 1 t o a given state 2
by two different processes, both of which are reversible, the integral of
6Q/T will be the same for both processes. More generally, we may say
that the
S
(6Q/T),,, depends only on the end states and not on the
intermediate series of states. The quantity 6Q/T for a reversible in-
finitesimal process is, therefore, an exact differential, and is the differ-
ential of a thermodynamic property. This property, which is called the
and this may be represented by the area under the curve representing
the process on a thermodynamic diagram of temperature versus entropy
(Fig. 2.2).
34 FOUNDATIONS OF THERMODYNAMICS Ch. 2
t Reversible
Adiabatic
Irreversible f'
Adiabatic
Irreversible
Adiabatic
FIG. 2.2. Reversible heat transfer on FIG. 2.3. Reversible and irreversible
temperature-entropy diagram. adiabatics.
where 6Q and T have the same meaning as in Eq. 2.9. This relation is
the most convenient analytical form of the Second Law.
When a system is isolated from all heat exchange with the surround-
ings, Eq. 2.13 shows that
(dS)adiabatic >
-0 (2.14)
which is the well known principle of the increase of entropy.
Since the entropy is defined by Eq. 2.11, it follows that the equality
sign necessarily applies in Eqs. 2.13 and 2.14 for reversible processes.
Furthermore, it may be shown that the inequality sign always applies
to irreversible processes.
Fig. 2.3 illustrates the paths of reversible and irreversible adiabatics
on the temperature-entropy diagram.
where 2, is the volume of the system and U is that part of the internal
Art. 2.4 THERMODYNAMIC PROPERTIES OF CONTINUUM 35
where e is the specific internal energy, or internal energy per unit mass,
E is the internal energy for the entire system, m is the mass of the sys-
tem, and so on. Small letters denote values per unit mass, and capital
letters represent values for the entire system. The specific volume is
the volume per unit mass and is, therefore, the inverse of the density,
where q, is the heat transfer per unit time per unit area in the x-direction,
and aT/ax is the temperature gradient in the x-direction. The minus
sign is inserted as the heat always flows from high to low temperatures.
In vector language, A is defined by
where q is the vector heat flux Der unit area. and vT is the vector gradient
of the temperature.
Art. 2.5 THE.FIRST LAW FOR A CONTROL VOLUME 37
Units and Dimensions. We shall adopt as standard units the ft, sec,
Ibf, slug, and degree Rankine. The units of some typical thermo-
dynamic quantities are then as follows:
Q f t ibf
u f t lbf/slug
c, ft lbf/slug R
s ft lbf/slug R
A lbf/sec R
a control volume, so now it is desirable to do the same for the two laws
of thermodynamics. In this article we shall do so for the First Law,
following identically the pattern of procedure employed in Chapter 1.
Consider the flow through the control volume of Fig. 2.4, with the
system defined as the material occupying the control volume a t time t.
We consider what happens during the time interval dt. Passing through
the control surface are a stationary strut and a rotating shaft attached
to a turbomachine, perhaps a compressor or turbine.
38 FOUNDATIONS OF THERMODYNAMICS Ch. 2
in terms of the rate of heat transfer, etc. Now we put the three terms
of this expression into more convenient form.
Rate of Heat Transfer. If 2 is the rate of heat transfer (ft lbf/sec)
through the control surface from the surroundings, then, as the interval
dt becomes vanishingly small, we may write
the rotating shaft resulting from the shear stresses in the plane cut by
the boundary of the system; (ii) the shear work done a t the boundaries of
the system on adjacent fluid which is in motion. These are respectively
denoted by and @&ear, the power delivered by means of a rotating
shaft and the power delivered by other shear forces. Note that shearing
forces a t the strut do not contribute to @shear,since the strut is not in
motion. Likewise, if the control surface coincides with the stationary
wall of a duct or machine casing, @shear is zero a t such points even though
there may be appreciable shear stresses. On the other hand, P8he,
plays an important role for a control surface lying within the viscous
boundary layer.
In the limit, as dt vanishes, Pahaft and @,hear of course refer t o the
material instantaneously occupying the control volume.
The First Law. Combining the expressions developed above, and
wing the expression for e given by Eq. 2.16 and the expression for h
given by Eq. 2.17, we finally obtain the following expression applicable
to the control volume:
where the integrals involve such terms as the "flux of enthalpy," "flux
of kinetic energy," etc.
We may rewrite Eq. 2.20 somewhat more compactly with the help of
vector notation:
and, if heat transfers across the control surface are negligible, there
results
82 >
sl (2.25)
Art. 2.7 THE PERFECT GAS
so that the change in internal energy between any pair of end states is
given by
42 FOUNDATIONS OF THERMODYNAMICS Ch. 2
Gas Constants. When a substance obeys the perfect gas law there is
a simple relation between c,, c,, and R. From the definitions of c,
and h, the perfect gas law, and the knowledge that h depends on T alone,
we form
dh d du d
- - (u 4- pv) = - + - RT = c, + R
dT dT
(2.29)
Of the four quantities c,, c,, R, and k, only two are algebraically
independent. With Eqs. 2.29 and 2.30, we may derive, for example,
k R
c, = -R ; c,=-- , etc.
k-1 k-1
For the most simple molecular model, the kinetic theory of gases shows
that
k=-n + 2
n
where n is the number of degrees of freedom of the molecule. Thus,
monatomic gases have n = 3, k = x. Diatomic gases such as oxygen,
nitrogen, etc., have n = 5 , k = 76. Extremely complex molecules,
such as freon and gaseous compounds of uranium have large values of
n and consequently values of k only slightly larger than unity.
Gas with Constant Specific Heats. The second part of the definition
of a perfect gas is that c,, which has already been shown to be dependent
on T only, is assumed to be constant. This restriction is not so good
an approximation to real gases as is the equation of state, and so some-
times we speak of a "semi-perfect gas," that is, one for which pv = RT
but having specific heats variable with temperature.
Art. 2.7 THE PERFECT GAS 43
Considering now a perfect gas, i t follows from Eq. 2.29 that .cp is also
a constant, and then Eqs. 2.27 and 2.28 yield
du pdv dT dv
&=--+-- -c,-+R-
T T T v
and, upon integration,
NOMENCLATURE
mean molecular velocity T absolute temperature
area V velocity
velocity of sound w mass rate of flow
specific heat at constant pressure W molecular weight
ratio of specific heats
mean free molecular path a Mach angle
length P bulk modulus of compression
Mach Number 6 boundary layer thickness
pressure X coefficientof thermal conductiv-
gas constant ity
universal gas constant P coefficient of viscosity
entropy per unit mass P density
- dV
Stotionory C
F I ~3.2.
. Pressure field produced by a point source of disturbance moving a t uni-
farm speed leftwards.
(a) Incompressible fluid (V/c = 0).
(b) Subsonic motion (V/c = x).
(c) Transonic motion (V/c = 1).
(d) Supersonic motion, illustrating Karman's three rules of supersonic flow
(Vlc = 2).
In each pattern the point 0 represents the present location of the dis-
turbance, the point - 1 represents the location one unit of time pre-
viously, and so on. For each of these previous locations there is drawn
a concentric circle showing the distance to which the corresponding wave
has spread. For example, to find the present location of the wave which
was emitted a t time -3, a circle is drawn with - 3 as a center and with
a radius 3 4 where t is the unit of time. The distance between point -3
and point 0 is then given by 3Vt where V is the velocity of the point
disturbance with respect t o the medium.
Art. 3.2 INCOMPRESSIBLE, SUBSONIC, AND SUPERSONIC FLOWS 51
passes over the ear of the observer; and why, when the latter does occur,
the noise is concentrated in a "crack."
Configurations like those of Fig. 3.2 may easily be observed in the
form of gravity waves on a free water surface when a sharp-pointed
-
object is drawn through the water a t varvine
speeds. The bow wave of a surface ship re-
" -
sembles the Mach cone of Fig. 3.2d.
Patterns like those of Fig. 3.2 can be made
visible in gas flows by means of shadow,
schlieren, or interferometer techniques, which
are optical methods for demonstrating density
variations in a gas. Fig. 3.3a, for example, is
a shadowgraph of a bullet which has just
passed through a cylindrical tube while travel-
ing a t supersonic speed. Fig. 3.3b is a schlie-
ren photograph of a similar arrangement, ex-
cept that slots are cut in the tube to allow
F I G. 3.4. Schlieren pho- only a selected number of wavelets to pass out
tograph of bullet traveling
at near the speed of sound of the tube into the field of view.
(after Aclceret). Fig. 3.4 shows a bullet traveling a t nearly
the speed of sound and having a wave front
similar to that of Fig. 3 . 2 ~ . At low subsonic speeds, corresponding to
Fig. 3.2b, no wave front appears.
Art. 3.2 INCOMPRESSIBLE, SUBSONIC, AND SUPERSONIC FLOWS 53
(b)
FIG.3.5. Cone-cylinder projectile traveling a t supersonic speed (Ballistic Research
Laboratory, Aberdeen).
(a) Schlieren photograph.
(b) Interferometer photograph.
54 INTRODUCTORY CONCEPTS TO COMPRESSIBLE FLOW Ch. 3
Let L be a characteristic length for the prototype, and let jrJ. be the
corresponding characteristic length for the model, where fL is the scale
factor for length. Similarly, we can summarize the other properties of
the prototype flow and model flow as follows:
Prototype Model
fvV
fop
~ P P
~ L L
fcc
fMM
fkk
Since the scale factors are constant, both of these equations can be satis-
fied only if
f p = fPfv2 (3.9a)
From Eq. 3.5 we obtain in similar fashion
c = dkp/p
and
so that
fcc = 4iix s
fc = m
Likewise, from Eq. 3.7, we have:
and
so that
or the Mach Number must be the same for the model and prototype if
the flows are to be similar.
When viscosity is present, a similar analysis applied to the inertia and
viscous terms in the Navier-Stokes equation leads to the criterion that
the Reynolds Number must be the same for similarity in flow patterns.
Dimensionless Groups Governing Compressible Flows. By consider-
ing all the physical equations which govern the flow, namely, the Navier-
Stokes equation, the continuity equation, the energy equation, and the
equation of state, i t is possible to arrive a t four dimensionless param-
eters which must be the same in order for two flow patterns to be sim-
ilar. These are
(i) The Mach Number
(ii) The Reynolds Number
(iii) The ratio of specific heats, k
(iv) The Prandtl Number, cPw/X,where c, denotes the specific heat
a t constant pressure, rc, the coefficient of viscosity, and X the
thermal conductivity.
POTENTIAL FLOW OUTSIDE BOUNDARY LAYER. For those regions of
the flow outside the boundary layer, where viscous effects and heat con-
duction effects are relatively unimportant, it is usually necessary that
only M and k be alike in order to have similarity. Of the two, similarity
in M is by far the more important, sincek has a relatively weak influence
on the flow pattern.
BOUNDARY-LAYER FLOW. In the boundary layer or in the interior of
shock waves, viscous and heat conduction effects are all-important, and
so the Reynolds Number and Prandtl Number must be included in the
similarity conditions. Fortunately, the Prandtl Number is nearly the
same for all gases and varies only slowly with temperature.
and
where a is the mean molecular velocity and I is the mean free molecular
path.
Using the foregoing relations, we now express the Reynolds Number
For this case Tsien (3) suggests that the realm of continuum gas dynam-
ics be limited to instances where the boundary-layer thickness is a t least
100 times the mean free path. That is,
58 INTRODUCTORY CONCEPTS TO COMPRESSIBLE FLOW Ch. 3
Fig. 3.7 shows the Reynolds Number per unit length as a function of
flight Mach Number for various altitudes, based on the standard atmos-
According to Eq. 3.11, the possible st,a,t.~sin the stream tube are rep-
resented in a diagram of c versus V by the steady-flow ellipse (Fig. 3.8).
Different parts of this ellipse represent schematically different realms of
compressible flow having significantly different physical characteristics.
Incompressible Flow. The velocity is small compared with the sonic
speed. Changes in c are very small compared with changes in V.
Subsonic Compressible Flow. The velocity and sonic speed a w of
comparable magnitudes, but the former is less than ithe latter. Changes
in Mach Number occur primarily Subsonic
because of changes in V, and only
secondarily through changes in c. Incorn-
press~ble '.,
C V.C Transonic
Transonic Flow. The di$erence
between V and c is small compared V>C
with either V or c. Changes in V v
"m,.
and c are of comparable magnitude.
FIG.3.8. Steady-flow adiabatic ellipse,
Supersonic Flow- The velocity showing realms of compressible flow.
and sonic speed are of comparable
magnitude,-but the former is larger than the latter. Changes in Mach
Number take place through substantial variations in both V and c.
Hypersonic Flow. The velocity is very large compared with the sonic
speed. Changes in velocity are very small, and thus variations in Mach
Number are almost exclusively the result of changes in c.
ma1 t o the paper and normal to the light beam. Thus, except for slight
refraction effects, each ray of light passing through the test section tra-
verses gas of constant density.
The chief difference between the two beams of light in Fig. 3.9 is that
the lower beam has passed through the test section. Since the different
rays of the lower beam are retarded by different amounts as they pass
through portions of the test section of different density, these rays will
COMPENSATOR
CHAMBER
!loUNT
TRANSLATOR
ROTATION MOUNT
COMPENSATING
HALF SILVERED
INTERFEROMETER
have various phase differences with the corresponding rays of the upper
beam when they are recombined, and thus an interference pattern will
be formed a t the screen.
where X and V denote respectively the wave length of the light and the
speed of light. This difference in traversal time is also rclatcd to the
difference in speed of light in the test section. Thus,
where L ik the length of test section along the light direction, and V b
and Va denote respectively the speeds of lisht in fluid b and a.
Art. 3.7 OPTICAL METHODS OF INVESTIGATION 63
-
Eliminatinp. t- "h - t.. from
-u -- ---- Ens.
-1-
3.12 and 3.13, and introducing Eqs. 3.14
and 3.15, we may obtain for t hie difference in index of refraction between
adjacent 1
na = X v m u o I L (3.16)
It remains to connect the index of refraction with the gas density.
This is done through the empirical Gladstone-Dale equation, which
states that
n-1
- --KG-D (3.17)
P
The right-hand side of this equation is easily computed from the di-
mension of the test section, the color of the monochromatic light em-
ployed, and the value of KG.D for air. Referring again t o Fig. 3.10,
the density in the low-speed flow upstream of t8hethroat is found by
measuring the temperature and pressure in this low-speed region. With
this as a reference, the density on each dark band in the nozzle may be
computed from Eq. 3.18, although the accuracy of this procedure using
the infinite-fringe interferogram is not high unless the optical compo-
nents are extraordinarily accurate.
FRINGE-DISPLACEMENT METHOD. For a more accurate quantitative
evaluation, the method described above is modified as now described.
Let us return to the situation where room air is in the test section of
Fig. 3.9. If the second splitting plate is rotated through a small angle
with respect to the first splitting plate, two "coherent" beams of light
which were in phase a t the first splitter will, through the change in
lengths of light paths, be out of phase a t the screen. If the splitters
are rotated about axes normal t o the paper, there will be formed on the
screen successive light and dark fringes, uniformly spaced, with each
fringe lying parallel to the axis of rotation (see, for example, Fig. 3.11a).
64 INTRODUCTORY CONCEPTS TO COMPRESSIBLE FLOW Ch. 3
where pz is the density in the test section, p l is the density a t the initial
reference condition, d is the distance between dark fringes in the refer-
ence condition, and I is the distance shifted by a dark fringe in passing
from condition 1 to condition 2.
When there is flow in the test section, with corresponding nonuniform
density changes, similar fringe shifts will occur, but, as they are no longer
uniform, the resultant fringes will be curved (Fig. 3.11b). Eq. 3.19 may
Art. 3.7 OPTICAL METHODS O F INVESTIGATION 65
then be applied a t each point in the flow. If both a flow and a no-flow
photograph are taken, Eq. 3.19 may be used to determine a t each point
the density change referred to the no-flow density. Or, from the flow
photograph alone, Eq. 3.19 may be used for finding the density differ-
ences between two points of the flow.
SUPERPOSITION METHOD. If the no-flow negative (Fig. 3.11a) is super-
imposed on the flow negative (Fig. 3.11b), the light passing through the
combination produces the light pattern of Fig. 3.11~. This is a particu-
larly convenient way of determining the contours of constant density,
for a consideration of the geometry of the two fringe patterns shows
that the fuzzy dark bands of Fig. 3 . 1 1 ~ are lines of constant fringe shift,
and, therefore, that they are lines of constant density. Indeed, the pat-
tern of Fig. 3 . 1 1 ~is like that of Fig. 3.10, and Eq. 3.18 may be applied
to either.
The Schlieren Method. Fig. 3.12 shows one of the numerous types of
schlieren (striae) arrangements. Light from a uniformly illuminated line
FL-I
Two- Dim.
f
Finite
Line Lens
source Screen or
Knife Photographic
Gloss Edge Plote
W0llS
z
Fra. 3.12. Schlieren system.
source of small but finite width is collimated by the first lens and then
passes through the test section. It is then brought to a focus by the
second lens and projected on the screen. At the focal point, where there
exists an image of the source, there is introduced a knife-edge which
cuts off part of the light. With no flow in the test section the knife-
edge is usually adjusted so as to intercept about half the light, and the
screen is uniformly illuminated by the portion of the light escaping the
knife-edge. When the flow is established in the test section (assumed
here, for simplicity, to be two-dimensional, with each light ray passing
through a path of constant air density) any light ray passing through a
region in which there is a density gradient normal to the light direction
will be deflected as though it had passed through a prism. Therefore,
depending on the orientation of the knife-edge with respect to the den-
sity gradirnt, and on t,he sign of the density gradient, more or less of
66 INTRODUCTORY CONCEPTS TO COMPRESSIBLE FLOW Ch. 3
the light passing through each part of the test section d l escape the
knifeedge and illuminate -the screen.
Thus the schlieren system makes density gradients visible in terms of
intensity cf illumination. A photographic plate at the viewing screen
records density gradients in the test section as different shades of gray.
THEORY OF LIGHT DEFLECTION. Let us assume that the flow in the
test section is parallel and in the x,y-plane, and that the light passes
through the test section in the z-direction.
Since the speed of a wave front of light varies inversely with the index
of refraction n of the medium through which the light travels, it follows
that a given wave front will rotate as it passes through a gradient in n.
Accordingly, the normal to the wave front will follow a curved path;
since this normal is what we mean by the light ray, the latter is refracted
as it passes through the density gradient. Noting that n is nearly unity
for gases, i t may be shown that
where R is the radius of curvature of the light ray. The total angular
deflection E of the ray in passing through the test section of width L is
therefore given by 7
Resolving this into Cartesian components, and taking note of Eq. 3.17,
we get for the angular deflections of the light in the x- and y-directions
FIG.3.13. Sehlieren photographs of flow through nozzle of Figs. 3.10 and 3.11
(M.I.T. Gas Turbine Laboratory).
(a) Knife-edge horizontal.
(b) Knife-edge vertical.
vln
vertical) the upper and lower halves look alike; the boundary layers are
not visible except on the inclined walls of the nozzle.
The Shadowgraph. A shadow system (Fig. 3.14) comprises simply a
small, bright source, a collimating lens, and a viewing screen or photo-
graphic plate. If the source is far
from the test section the collimat-
ing lens is unnecessary.
4
Assume a t first that the test sec- w l n t
Source
tion has stagnant air in it and that
Viewing
the intensity of illumination on the Flow Screen or
Photographic
screen is uniform. When flow is Piate
established in the test section the F IG. 3.14. Shadowgraph system.
light beam will be refracted wher-
ever there is a density gradient. However, if thc- density gradient were
constant, every light ray would be deflected by the same amount, and
there would be no change in illumination on the screen. Only if there
is a gradient in density gradient will there be any tendency for the light
rays to diverge or converge. From this it is evident that variations in
illumination of the screen are proportional to the second derivative of
the density gradient, i.e., to the term
aZP aZp
-+-
ax2 ay2
assuming two-dimensional flow in the x,y-plane.
68 INTRODUCTORY CONCEPTS TO COMPRESSIBLE FLOW Ch. 3
(a) (b)
FIG.3.15. Detached shock in front of blunt body (Ordnance Aerophysics Labora-
tory).
(a) Shadowgraph.
(b) Schlieren, knife-edge horizontal.
side. In the shadow picture the shock therefore shows as a dark line
followed by a bright line. For comparison, Fig. 3.15b shows a schlieren
photograph of a similar shock.
R E F E R E N C E S A N D S E L E C T E D BIBLIOGRAPW
1. VON KLRMLN,T H. Supersonic Aer~dy~nlics-l'rinciplcs and Applications,
Jour. Aero. Sci., Val. 14, No. 7 (1947), p. 373.
2. JEANS, SIR JAMES. An Introduction to the Kinetic Theory of Gases. Cambridge:
Cambridge University Press, 1946.
3. TSIEN, H. S. Superaerodynamics, Mechanics of Rarefied Gases, Jour. Aero. Sci.,
Vol. 13, No. 2 (1946), p. 653.
4. SCHARDIN, H. Dm Toeplersche Schlierenverfahren, Forschungsheft V.D.I., 367,
Ausgabe B, Band 5 (.July-August, 1934).
5. ZOBEL, TH. Development and Construction of an Interferometer for Optical
Measurements of Density Fields, NACA Tech. Memo., No. 1184 (1947).
6. B ARNES, N. F., and B ELLINGER , S. L. Schlieren and Shadowgraph Equipment
for Air Flow Analysis, Jour. Opt. Soc. Amer., Val. 35, No. 8 (1945), p. 497.
7. L ADENBURG , R., W INCKLER, J., and V AN VOORHIS,C. C. Interferametric
Studies of Faster Than Sound Phenomena, Phys. Rev., Vol. 73, No. 11 (1948),
p. 1359.
8. ASHKENAS, H. I., and BRYSON, A. E. Design and Performance of a Simple Inter-
ferometer for Wind-Tunnel Measurements, Jour. Aero. Sci., Val. 18, No. 2
(1951). n 82.
Art. 3.7 PROBLEMS 69
9. DEFRATE, L. A., BARRY, F. W., and BAILEY, D. Z. A Portable Mach-Zehnder
Interferometer, Meteor Report, No. 51, M . I. T., Cambridge, Mass. (1950).
10. SCHARDIN, H. Theorie und Anwendung des Mach-Zehnderschen Interferenz-
Refraktometers, Zeilschrift jur Instrumentenkunde, Vol. 53 (1933), pp. 396,
424. Also see R.A.E. Translation, No. 79.
11. LIEPMANN, H. W.,and PUCKEIT, A. E. Introduction to Aerodynamics of a Com-
pressible Fluid. New York: John Wiley & Sons, Inc., 1947.
12. PANKHURST, R. C., and HOLDER, D. W. Wind- Tunnel Technique. London:
Sir Isaac Pitman & Sons, 1952.
PROBLEMS
3.1. The compressibility of a liquid is usually expressed in terms of the bulk
modulus of compression,
8 = P-d p
dp
Show that
Stagnation
Sectipn Confrol
////,/,/,,/,
To 1 w'-C Flow
Ti
I
/ / / / / I / / /
, /
FIG. 4.1. Flow between stagnation sec- FIG. 4.2. Isentropic ac-
tion and any other section. celeration or deceleration on
Mollier chart, showing stag-
nation enthalpy and isen-
tropic stagnation pressure.
increase in cross section. The pressure ratio, p/po, where the flow per
unit area is a maximum, is called the critical pressure ratio, and has a
value, for all real gases and vapors, of approximately one-half.
Pressure ratios greater than the critical correspond to subsonic flow,
and pressure ratios less than the critical correspond to supersonic flow.
This will now be demonstrated in a completely general manner.
Distinction Between Subsonic and Supersonic Flow. We first write
the steady-flow energy equation in differential form for two cross sec-
tions infinitesimally distant from each other. Thus
so that
giving
Art. 4.2 GENERAL FEATURES OF ISENTROPIC FLOW 77
Substituting Eq. 4 . 2 ~into E q . 4.2d and rearranging, we get
~ P D P= (apiap), = c2
so that
-- 1 - M~
dp (4.2e)
pV2
Now, from Eq. 4.2c, which is the dynamic equation for a frictionles:.
fluid, it is seen that the pressure always decreases in an accelerating flow.
and increases in a decelerating flow. In other words,
Using this result in conjunction with Eq. 4.2e, we arrive a t the follow-
ing conclusions of practical significance:
(i) For subsonicspeeds (M < l),
Thus, we have the astonishing result that the effects of an area change,
say an increase in area, are exactly opposite for subsonic and supersonic
flow.
The possible types of flow, according to this tabulation, are summa-
rized schematically in Fig. 4.4.
A t Mach Number unity the area goes through a minimum. This
important conclusion is valid irrespective of the type of fluid considered,
whether gaseous or liquid.
In constructing the curves of Fig. 4.3, the equation of state, Eq. 4.ld,
was employed. Since the equation of state for real gases and vapors can
seldom be put into simple algebraic form, the curves of Fig. 4.3 cannot,
in general, be formulated analytically, but instead are found through
78 ISENTROPIC FLOW Ch. 4
direct computation. If, on the other hand, it is assumed that the perfect
gas laws are valid, analytical results are obtainable, and the numerical
computation of problems is greatly simplified. For many engineering
gases, particularly air a t moderate pressures and temperatures, the
deviations from the perfect gas laws are negligible; hence, most calcula-
tions are based on these simple relations.
(M>I1 V increases
SUBSONIC SUPERSONIC
O Z Z LE
\
-Flow
-
Flow
-
p decreases p increases
(M<I) V increases V decreases
/--
- T H ROAT
Flow
( M I1
dp50
dMe0
FIG. 4.4. Effects of area change on pressure and velocity in subsonic and super-
sonic flow.
Using Eq. 4.3a and 4.3d, the steady-flow energy equation (Eq. 4.la)
becomes
2k
V = 42/2c,(~o- T ) = d/--R(To - T) (4.4)
k-1
From this we see that for a fixed stagnation temperature To, (sometimes
called total temperature), all states with the same temperature have the
same velocity. Referring to the temperature-entropy diagram of Fig.
4.5, lines of constant velocity are horizontal, and the vertical distance
between To and T is proportional to the square of the velocity.
l rreversible
T Deceleration
T*
I C I
S S
(4 (b)
FIG.4.5. Flow processes on temperature-entropy diagram.
(a) Isentropic acceleration or deceleration.
(b) Irreversible adiabatic deceleration.
(4.5a)
or, using Eq. 4.4 and the equation for the sound speed in a perfect gas,
fir-
Now, using Eqs. 4.5, we get the following relations between the three
reference velocities, together with the numerical values for k = 1.4:
Since cp = kR/(k - 1) and cZ = kRT, this takes the simple and con-
Note that Eqs. 4.5 through 4.7 may be obtained rather easily from
Eq. 4.8.
The Dimensionless Velocity M*. As a dimensionless parameter the
Mach Number is very convenient, but it has two disadvantages: (i) i t
is not proportional to the velocity alone and, (ii) a t high speeds it tends
towards infinity. Often, therefore, i t is useful to work with a dimension-
less velocity obtained through dividing the flow velocity V by one of the
three reference velocities of Eq. 4.8. Generally the most useful of these
ratios is defined by
M* = V/c* = V/V*
It should be noted immediately that although in general the asterisk
denotes the value of a property a t Mach Number unity, this convention
is not followed in the definition of M*. The latter is not the value of M
at the local sonic condition, but is rather defined as given above.
There is a unique relation between M and M* for adiabatic flow. From
the definitions of M* and M,
Furthermore, the first and last parts of Eq. 4.8 may be divided by c*' to
give
L
Eliminating c2/c*' from this, pair of equations, and rearranging, we get
the useful formulas
and
82 ISENTROPIC FLOW Ch. i
Thus, the temperature a t the throat is only about 17 per cent less
than the stagnation temperature, whereas the throat pressure is only
about half the isentropic stagnation pressure.
The critical pressure ratio, p*/po, is of the same order of magnitude
for all gases. It varies almost linearly with k from 0.6065 for k = 1 to
0.4867 for k = 1.67.
Mass Flow Relations in Terms of Mach Number. From Eqs. 4.14
and 4.10 it is clear that either T/To, p/po, M or M* may be taken
as an independent parameter, and that the remaining quantities would
then depend uniquely on the value of the chosen independent parameter.
By and large, the variable M has been found to be the most convenient
choice as far as simplicity of practical calculations is concerned. We
shall, therefore, follow the practice in this and succeeding chapters of
deriving all the working formulas in terms of M as the independent
variable.
To find a convenient formula for the mass flow per unit area in terms
of M, we eliminate p in the equation preceding Eq. 4.11 by means of the
isentropic law (Eq. 4.14b). Thus we obtain
This shows that, for a given Mach Number, the flow is proportional to
the stagnation pressure and inversely proportional t o the square root
of the stagnation temperature. For this reason, flow test data on com-
pressors and turbines, or indeed on any flow passage which operates
over a wide range of pressure and temperature levels, are usually plot-
ted with w f i / p o as the flow variable. In this way the results of a
given test become applicable to operation a t levels of temperature and
Dressure different from the original test conditions.
Art. 4.4 ISENTROPIC FLOW OF A PERFECT GAS 89
Now it is evident that if M were eliminated from Eq. 4.16 with the
help of Eq. 4.14b, we would have w/A in terms of k, R, po, To, and p/po.
The resulting expression is then the algebraic formula, for a perfect gas,
corresponding to the curve of w/A versus p/po in Fig. 4.3.
MAXIMUM FLOW PER UNIT AREA. T O find the condition of maximum
flow per unit area, we could compute the derivative d(w/A)/d(p/po)
and set this derivative equal to zero. From this we would find that the
critical pressure ratio is given by Eq. 4.15b.
An equivalent procedure would be to differentiate Eq. 4.16 with re-
spect to M and set this derivative equal to zero. At this condition, we
would find that M = 1.
However, neither of these procedures is necessary inasmuch as we
have proved quite generally in Art. 4.2 that the cross-sectional area for
isentropic flow passes through a minimum a t Mach Number unity.
Therefore, to find (w/A),,,, we need only set M = 1 in Eq. 4.16.
Thus we find
For a given gas, therefore, the maximum flow per unit area depends
only on the ratio For given values of the stagnation pressure
and stagnation temperature and for a passage with a given minimum
area, Eq. 4.17 shows that the maximum flow which can be passed is
relatively large for gases of high molecular weight and relatively small
for gases of low molecular weight. Doubling the pressure level doubles
the maximum flow, whereas doubling the absolute temperature level
reduces the maximum flow by about 29 per cent.
FLIEGNER'S
FORMULA. Using the values k = 1.4 and R = 53.3
O
ft lbf/lbm R, corresponding to air, we obtain from Eq. 4.17
The area ratio is always greater than unity, and for any given value of
A/A* there always correspond two values of M-one for subsonic flow,
and the other for supersonic flow.
The Impulse Function. For problems involving jet propulsion it is
sometimes convenient to employ a quantity called the impulse function,
defined by
F = pA pAV2+ (4.20)
Applying the momentum equation to the flow through the control sur-
face of Fig. 4.7, i t is seen that
and so
substituting p/po, p ~ / p * ,and A/A* from Eqs. 4.14b, 4.15b, and 4.19,
respectively, there is obtained after simplification,
Next, we find the properties at the throat by selecting values from Table
B.2 at M = 1:
p*/po = 0.528; :. p* = lO(0.528) = 5.28 psia
Alternatively, the mass flow may be computed from the continuity equation
at the throat or test section. For example,
.%I%. 4.6 CHOKING I N ISENTROPIC FLOW
fvariablel
p,= conFt
Exhauster
Valve
Regime
0
Distance A l o n g Nozzle
(4
0 PB'Po I 0 Pe'Po I
(b) (c)
Fro. 4.11. Operation of converging nozzle at various back pressures.
Regime I Regime I1
w 6
dependent on pB/po
wd?& independent of p ~ / p ,
AEPO AEPO
Art. 4.7 OPERATION O F NOZZLES 93
this case p is the static pressure and p, is the pressure measured a t the
mouth of the tube. If V is in error by 1 per cent, then V2 is in error by
2 per cent. Hence we set
( =0.02; v
o m which - r 0.28
4 co co
The latter figure corresponds to an air speed a t normal temperatures of
about 300 ft/sec. At higher speeds the error increases quite rapidly.
Isentropic Formulas in Powers of Mach Number. Expanding Eqs.
4.9, 4.14, 4.16, and 4.19 in powers of M2 by means of the binomial the-
orem, the following convenient formulas for low-speed isentropic flow,
valid up to orders of M4, may be found:
where 0 = 5526OR for air. The results are embodied in Fig. 4.13, in
which are shown the per cent errors in the pressure, temperature, and
density ratios incurred through use of constant rather than variable
specific heats. At stagnation tem-
peratures of 1000°R or less, the -16
error is seen to be small for en-
gineering purposes; but, a t tem- -12
peratures greater than 2000°R, the
error can be substantial, especially
a t supersonic Mach Numbers. 2
-8
..
-
L
W
-2 -4
"..
C
a
L m
0 -I
L
a
0
-
W
C
Ql
U 0
a,
4
a
1 8
0 2 4 6 8 10 0 2 4 6 8 10
M M
(4 (d)
FIG.4.13. Error incurred through assumption of constant specific heats for air,
nith p = pRT (after Donaldson).
Fro. 4.14. Effect of pressure level on isentropic flow functions, using van der Waals
equation for air (after Donaldson).
FIG. 4.15. Simultaneous effects of high prcssure and high temperature on pressure
ratio and area ratio (after Donaldson).
98 ISENTROPIC FLOW Ch. 4
- ve
Pe
passage. Consider a nozzle (Fig.
4.16) supplied with gas a t low
velocity and a t pressure pi and
temperature Ti. The gas expands
adiabatically, but with increasing
T~ entropy, to state e. If it had ex-
T
panded without friction to the
same final pressure, the end state
would have been s. We now define
Tc
the nozzle efficiency as the ratio of
the exit kinetic energy to the exit
Pe
kinetic energy which would be ob-
--
S
tained in a frictionless nozzle ex-
Fro. 4.16. Illustrating definition of
nozzle efficiency. panding the gas to the same final
pressure. With the help of the
steady-flow energy equation the efficiency may be written
which is convenient for reckoning the exit velocity when the efficiency,
pressure ratio, and stagnation temperature are all known.
Art. 4.10 PERFORMANCE OF REAL NOZZLES 99
w
C, =
Isentropic Flow Rate
If the over-all pressure ratio of the nozzle is such that the velocity a t
the minimum section is subsonic, then the "isentropic flow" is reckoned
in terms of the exit conditions of the nozzle. However, if the pressure
ratio is such that sonic velocity prevails a t the minimum section, then
the "isentropic flow" is reckoned by using the formula for choking flow
a t the throat. These specifications apply to both converging and con-
verging-diverging passages.
The remarks made previously concerning the factors influencing
nozzle efficiency pertain also to the discharge coefficient.
For well-designed nozzles with straight axes having "pipe" Reynolds
Numbers measured at the minimum area of lo6 or more, the discharge
coefficient is of the order of 0.99, but it may be considerably less for
low Reynolds Numbers.
Neither the discharge coefficient nor the velocity coefficient of rounded-
entrance nozzles suitably designed for the operating pressure ratio are
significantly dependent upon the leaving Mach Number.
100 ISENTROPIC FLOW Ch. 4
-
charge coefficient for a sharp-edged orifice meter is due primarily to the
contraction ( v e n a contracts) in the
/ / / / / / / / / / / / / / / / stream following the orifice. The
PI "2 contraction in turn is due to three-
//////,,/,,,/// dimensional effects. The coefficient
of contraction increases substan-
tially as the result of compressi-
0.9
bility effects (Fig. 4.17). It should
0.8
be noted that the isentropic flow
Cw
on which the discharge coefficient
0.7 is based is reckoned as though a
rounded-entrance converging noz-
0.6 zle, having the same exit area as
I 0.6 0.4 0.2 O the orifice, were supplied with gas
PZ /PI a t stagnation pressure pl and dis-
FIG. 4.17. Discharge coefficient of h d to a region having the
sharp-edged orifice meters with zero ve- c arge
locity of approach (after J. A. Perry). pressure pz.
Since rocket motors generate gas a t about 500 psia and operate in
atmospheres a t 14.7 psia or less, a converging-diverging nozzle is usually
used. Except under operating conditions far from the design point,
sonic conditions occur at the throat, and the flow to the nozzle exit is
shock-free. We shall assume these conditions for the present analysis.
Art. 4.1 1 SOME APPLICATIONS O F ISENTROPIC FLOW 101
Since the pressure ratio p,/po depends only on the area ratio, Eq. 4.33
indicates that the thrust for a nozzle of given size and geometry depends
only on po and the ratio pa/po, and is independent of the temperature To.
EFFECT O F A RE A RATIO. We now ask, for given values of At, pol and
p,, what exit area should be used in order to obtain maximum thrust?
By applying the calculus to Eq. 4.33 it may be shown after a laborious
calculation that 3 is a maximum when the area ratio is chosen i n such a
way as to make the pressure in the exit plane exactly equal to pa. How-
102 ISENTROPIC FLOW Ch.4
ever, this result may more easily be obtained with the simplest of phys-
ical reasoning. The net thrust on the rocket is the resultant of static
preseures acting on all the surfaces of the motor. Suppose, as in Fig.
4.18b, that there is a certain exit area for which p, = pa. If the nozzle
is continued beyond this point, the pressure in the nozzle will drop
further, and the added piece of divergent nozzle will have negative thrust
because the internal pressure on this added piece is less than the external
pressure. By similar reasoning, i t follows that cutting off a piece of
nozzle upstream of the plane where p, = p, would also act to reduce
the thrust. Hence we conclude that the thrust is a maximum when
p, = pa. Applying this criterion to Eq. 4.33; we get
If the nozzle were a simple converging nozzle, A, would equal At, and
p,/po would be the critical pressure ratio. Making these substitutions
Art. 4.1 1 SOME APPLICATIONS OF ISENTROPIC FLOW 103
To illustrate the effect of area ratio on nozzle thrust, we form the ratio
--
( 2 ) (l - -A)++ -& ( p e- - -Pa)
k+l
k-1 k + l At Po
- (4.36)
This ratio is plotted against area ratio in Fig. 4.19 for various values
of p,/po, with k = 1.2 (typical value for rocket gas). It is seen that the
curves are quite flat near their maxima, so that the area ratio need not
be exactly adjusted in order to obtain substantially maximum thrust. I n
practice, rocket nozzles are usually designed with p, greater than pa,
since this reduces the size of the nozzle without materially reducing the
thrust.
Reynolds Number for Supersonic Wind Tunnel. Fig. 4.20 shows a
supersonic wind tunnel with a test section having a Mach Number MI
With Friction
FIG. 4.21. Reynolds Number for su- FIG. 4.22. Flow in duct with friction.
personic wind tunnel (NACA Tech. Note,
No. 1428).
pressures pi and p, together with the discharge coefficient C,, the local
Mach Number M may be found from these tables by a quick computa-
tion. An illustrative example is given in Chapter 6.
Eq. 4.38 typifies a technique which we shall find useful from time t o
time-namely, t o use the tabulated functions of k and M in Appendix B
not only for the particular types of flow underlyi~gthe construction of
the tables, but for such other problems where identical functions of k
and M appear.
PROBLEMS
4.1. Consider the reversible adiabatic flow of steam through a passage of
variable cross section. At the section where the velocity is zero the pressure
and temperature are po = 50 psia and TO = 800°F. Denoting the pressure a t
any other point in the stream by p, plot against p/po the values of specific vol-
ume (ft3/lb), velocity (ft/sec), and mass velocity (lb/ft2 sec), for the following
conditions:
(a) The properties of steam are taken from the Steam Tables of Keenan
and Keyes.
(b) The steam is considered as a perfect gas, with a value of 1.3 for k.
(c) The steam is considered as incompressible with a density equal to the
density corresponding to po and To.
In the above calculations choose for the lowest value of p/po the value of p
which corresponds to the first appearance of moisture in part (a).
4.2. -consider the reversible, adiabatic flow of a perfect gas. Plot the values
of p*/po, T*/To, p*/po, c*/co, Vm,/c*, and Vm,/co, all versus k, for values of
the latter between 1 and 2.
4.3. An airplane flies at an altitude of 40,000 ft (temperature = -67.0°F,
pressure = 2.72 psia) with a speed of 400 mph. Neglecting frictional effects,
(a) Calculate the critical velocity of the air relative to the aircraft.
(b) Calculate the maximum possible velocity of the air relative to the
aircraft.
4.4. Sketch a curve of pressure (p/p*) versus velocity (V/V*) for isentropic
flow, paying special attention to zero or infinite slopes, direction of curvature,
and points of inflection. Indicate the values of p/p* and V/V* a t their maxi-
mum and minimum points, at points of zero or infinite slope, and a t points of
inflection. What is the physical significance of the tangent to the curve of p/p*
versus V/V*?
106 ISENTROPIC FLOW Ch. 4
4.5. A stream of air flows in a duct of 4 inches diameter a t a rate of 2.20
Ib/sec. The stagnation temperature is 100°F. At one section of the duct the
static pressure is 6 psia.
Calculate the Mach Number, velocity, and stagnation pressure a t this section.
4.6. A perfect gas (k = 1.4, R = 100 f t lbf/lbmOR) is supplied to a converg-
ing nozzle a t low velocity and a t 100 psia and 540°F. The nozzle discharges
to atmospheric pressure, 14.7 psia. Assuming frictionless adiabatic flow, and a
mass rate of flow of 1 lbm/sec, calculate
(a) The pressure in the exit plane, in psia
(b) The velocity in the exit plane, in ft/sec
(c) The cross-sectional area of the exit plane, in square feet
4.7. Show that for isentropic flow of a perfect gas, the pressure, temperature,
and density, when made dimensionless with respect to the corresponding critical
values, are given by
(a) Show that the value of the pressure coefficient corresponding to the
appearance of the critical velocity is given by
(b) Plot log (-Cp*) versus log M , for k = 1.4 and for values of M, between
0.1 and 1.0.
(c) Suppose that an airplane is flying a t sea level with a velocity of 500 mph.
What is the maximum pressure coefficient which may be attained on the wings
without the speed becoming anywhere supersonic?
4.15. Pressure coefficients, lift coefficients, drag coefficients, etc., of airfoils
which are in a free stream with conditions p,, &, etc., are usually expressed in
terms of the dynamic head of the free stream. Thus
(b) Show that the field of flow outside the minimum radius includes all Mach
Numbers from zero to infinity, and that the radius corresponding to the critical
velocity is given by
4.20. Derive the following expressions for isentropic flow with the pressure
ratio p/po as a parameter:
V = pk -%l E d =
4.21. Derive relations between M and V/co and between M and V/V-,
for adiabatic flow of a perfect gas.
4.23. By expanding Eq. 4.11 in a power series of M with the aid of the
binomial theorem, show that for low Mach Numbers the mass flow parameter
may be approximated by
4.24. Derive Eqs. 4.14 without use of the steady-flow energy equation, by
employing Euler's equation for frictionless flow, dp = -pV dV, and the perfect
gas relations p = pRT, c2 = kRT, and plpk = constant.
(c) Show that Mach Number unity does not occur at the minimum area, and
find the Mach Number a t the throat.
4.27. Consider a supersonic nozzle constructed with a ratio of exit to throat
area of 2.0. The nozzle is supplied with air at low speed at 100 psia and 140°F.
The over-all nozzle efficiency from inlet to exit is 90 per cent, but the flow is
isentropic .to the throat.
Calculate the pressure, velocity, and Mach Number at exit, and compare
with. the corresponding values for isentropic flow.
4.28. A tank having a volume of 100 ft3 is initially filled with air at 100 psia
and 140°F. Suddenly the air is allowed to escape to the atmosphere (14.7 psia)
through a frictionless converging nozzle of one-inch diameter. It is agreed to
assume that the flow is quasi-steady, i.e., that the steady flow equations may be
applied to the nozzle at any instant of time. Furthermore, the tank is to be
considered as insulated perfectly against heat conduction and as having no
heat capacity.
Plot the pressure in the tank versus elapsed time.
4.29. A supersonic nozzle with a throat area of 1 sq in. discharges air into a
duct having a cross-sectional area of 2 sq in. The supply pressure is 100 psia,
and the nozzle has a discharge coefficient of 0.98. At one section of the duct
the pressure is 14.2 psia. Calculate the Mach Number and isentropic stag-
nation pressure at this section.
Ch. 4 PROBLEMS 111
where 8 is a constant.
8= r (g)
(a) Show that
pa - p* = @ I n 2
where po is the stagnation pressure and p* is the critical pressure.
(b) Derive expressions for p / p ~and A/A* in terms of M and 8.
4.31. A large main is connected to an evacuated tank with a volume of 10
3
ft by means of a rounded-entrance,
/'//''''/'''/
- ,,E
converging nozzle having a diameter
of 0.01 in. Initially, a diaphragm " ' / / / / ' / '
over the orifice seals the tank from Main rank
the main. The air in the main is a t ,,,////////,,
100 psia and 70°F. The diaphragm
,
3 ,,/,/,///////,/
is suddenly broken and air rushes P ROB. 4.31.
into the tank. Estimate the time
required for the pressure in the tank to reach 25 psia, based on the following
assumptions:
(i) The flow is quasi-static.
(ii) There is no heat conduction from the tank to the air.
(iii) The pressure and temperature in the main are constant.
Under what circumstances will these assumptions lead to accurate results?
4.32. Show that the coefficient of contraction for a Borda re-entrant orifice
only discontinuities normal to the direction of flow (as i n Fig. 5.1). Ob-
lique discontinuities will be treated i n C h a p t e r 16.
Additional material relevant to the s u b j e c t m a t t e r of the present chap-
ter may be found i n Volume 11, C h a p t e r s 25 and 28.
NOMENCLATURE
h,+?=
vz2 v,2
h v t Z - - ho (5.1)
ous values of V,, the Fanno line may easily be constructed. Since the
momentum equation has not yet been introduced, the Fanno line rep-
resents states with the same flow per unit area and the same stagnation
enthalpy, but not necessarily with the same value of the impulse func-
tion. Thus, in general, frictional effects are required to pass continu-
ously along the Fanno line from state x to any other state on the line.
DdV+ Vdp = 0
116 NORMAL SHOCK WAVES Ch. 5
From the continuity expression, Eq. 5.2, and the relation p = pRT,
Combining Eqs. 5.11 and 5.12, we find the equation of the Fanno line
in terms of p and M,
Finally, after eliminating pulpz from Eqs. 5.13 and 5.14, we get a
<
The first of these solutions is trivial since it expresses the obvious fact
that conditions may be identical a t sections x and y. Inspection of Eqs.
5.11 and 5.13 shows that, for the trivial solution, p, = p, and T, = T,.
The second solution indicates the connection which must subsist be-
tween the states on the two sides of the discontinuity. In the subse-
quent discussion attention is given to the discontinuity solution.
Substituting for M, and M, from Eq. 4.10 gives the following simple
relation between M,* and M,* known as Prandtl's equation:
M,*M,* = 1
or, since c,* = c,*,
v,v, = c*~
An alternative form is
Substituting the value of M; from Eq. 5.16b into Eq. 5.11, we get
for the temperature ratio,
k - 1 ME2) (- 2k M,' - 1)
Tu -
- - k-1 (5.19)
T* (k +
112
MZ2
2(k - 1)
Art. 5.3 NORMAL SHOCK IN A PERFECT GAS 119
The density ratio may be found in terms of Mz from Eqs. 5.18a and
5.19, and the perfect gas law,
Po,- Po, P, Pz
-
P O ~ P, Pz P O ~
Now p,/p, is given by Eq. 5.18a, and po,/p, and pZ/poz may b e found
from Eq. 4.14b. Using Eq. 5.16b for the value of M,, we get, after
algebraic simplification,
Also, since
and
FIG.5.4. Entropy change across nor- FIG. 5.5. Illustrating the impossibility
mal of a rarefaction shock.
is less than unity. The general form of Eq. 5.24 is shown graphically
in Fig. 5.4. It is thus proven rigorously that for a perfect gas only the
shock from supersonic to subsonic speed is possible.
The irreversibility of the normal shock for perfect gases can also be
demonstrated with the help of the temperature-entropy diagram, Fig.
5.5. Since, for a perfect gas, the enthalpy is a linear function of tem-
perature, the Fanno and Rayleigh curves of Fig. 5.5 are similar to the
corresponding curves of Fig. 5.3. Continuous processes which follow
Art. 5.3 NORMAL SHOCK I N A PERFECT GAS 121
the Rayleigh line involve reversible heat transfer. Imagine that state y
is reached reversibly from state x by adding heat a t constant area along
the Rayleigh line from x to b and then rejecting heat from b to y. Since
the process is reversible, the area under the curve on the T-s diagram
represents the net heat transfer, i.e.,
Now, since the stagnation temperatures are the same a t x and y, the net
heat transfer for the process considered must be zero, and, consequently,
the net area under the curve xby must be zero. The heating from x to b
is at low temperatures, and the cooling from b to y is a t high tempera-
tures. Hence state y must have a greater entropy than state x.
Since s, must exceed s,, the equation preceding Eq. 5.24 leads to the
important conclusion that there is alvays a decrease in isentropic stag-
nation pressure across a shock wave.
The Rankine-Hugoniot Relation. An interesting relation connecting
the pressure ratio and density ratio is known as the Rankine-Hugoniot
equation. Substituting the value of M; from Eq. 5.16b into Eq. 5.11,
we obtain, after algebraic rearrangement,
Pz
-+--kk -+1l
P,
PZ
--
Pz k+l P,
-
This relation is plotted in Fig. 5.6, together with the isentropic rela-
tion between pressure and density (p pk). It is seen that for diatomic
gases the density can at most increase by a factor of six, whereas the
122 NORMAL SHOCK WAVES Ch. 5
pressure ratio may reach infinity. The curves indicate another impor-
tant feature, namely that weak shocks are nearly isentropic; by "weak
shock" we mean a shock in which the percentage rise in pressure is small.
py4 Mx
F I G . 5.6. The Rankine-Hugoniot FIG.5.7. Working curves for normal
mrve, for k = 1.4. shock, for k = 1.4.
Fig. 5.7 gives the important shock relations in graphical form for a
gas with k = 1.4.
To facilitate extensive or accurate numerical calculations, the normal
shock functions are tabulated in Table B.3 for k = 1.4.
(E) = 0.1278; .
I
10
p , = -- 78.4 psia
0.1278
Also, since
c, = 4 9 . 1 =~49.1- = 1087 ft/sec
it follows that
V, = M&, = 2(1087) = 2174 ft/sec
Now, from Table B.3, at M, = 2.00,
V,/Vv = 2.667; p@/po, = 0.7209
whence
Vv = 2174/2.667 = 815 ft/sec
p~ = (78.4)(0.7209) = 56.5 psia
From Eq. 5.29 it appears that the term ( ~ -21) is directly propor-
tional to the shock strength. If the shock is relatively weak, Eq. 5.30
indicates that the ratio of density increase to initial density,
(5.31;
Now, for weak shocks, each of the terms on the right-hand side of
Eq. 5.31 is of the form
In (1 +
E)
where E is much smaller than unity. We therefore employ for each term
the series exvansion
Carrying out these operations and simplifying, we find that the first-
order and second-order terms in the shock strength vanish, leaving
In terms of (MZ2- l), we find by substituting Eq. 5.29 into Eq. 5.32b
that the entropy increase is given by
Thus, for weak shocks, the entropy increase is of the third order with
respect to the shock strength. As a first approximation, therefore, the
irreversibility connected with weak shocks may be ignored, and the
isentropic relation between pressure and density may be used for con-
necting the states on both sides of the shock.
This result explains why in Fig. 5.6 the isentropic curve and Rankine-
Hugoniot curve have the same slope and the same radius of curvature
a t the lower left-hand corner of the chart. It explains also why in Fig.
5.7 there is scarcely any loss in stagnation pressure across the shock
for initial Mach Numbers less than about 1.25.
:
t=2
Therefore, in order to explain the existence of
shock waves in a duct or attached to a moving
m:4;
a : .
body, i t is necessary to examine the way in
which the steady-state condition is achieved
from a condition of no motion.
ph; :. !:
Physical Description of Wave Development.
The simplest physical process incorporating the
elementary ideas expressed above is illustrated
p k X * in Fig. 5.8. The piston accelerates rightward
i . e,f into gas which is initially motionless in the
0 ib:=:di
I
duct. For purposes of analysis we will sup-
FIG. 5.8. Wave fronts
formedat successive times Pose that the gradual acceleration is approxi-
as piston accelerates mated by a series of instantaneous increases of
rightward by a series of
spaced impulses. speed occurring a t equal intervals of time, with
the piston traveling a t uniform speed between
each pair of impulsive accelerations. Each successive pressure wave
produced follows the laws derived in Art. 3.2.
In Fig. 5.8 the pressure distribution in the duct is shown schematically
at various times after the beginning of motion. The position of the
vertical axis on each chart indicates the instantaneous location of the
piston. After the first impulsive acceleration of the piston, i.e., a t t = 1,
the pressure wave (shown as a vertical front) has moved down the duct
a short distance and influenced the mass of gas labeled "a." The gas
mass "a" is therefore a t a slightly increased pressure and is moving to
the right with the velocity of the piston. Between t = 1 and t = 2,
the mass l Ia ,7 receives a further increase in pressure and velocity from
the second acceleration of the piston, and the original wave travels
along the duct and imparts a pressure and velocity pulse to mass 'lb."
This process continues as each of the pressure waves travels downstream
and as fresh waves are initiated by successive piston accelerations.
Art. 5.6 FORMATION OF SHOCK WAVES 1
3
Now each of these pressure waves travels with the local speed of sound
relative to the fluid through which it is passing. But the masses nearer
the piston have forward motions greater than those further from the
piston. Moreover, since the process is isentropic, the masses nearer
the piston have greater sound velocities by virtue of the higher tempera-
tures associated with their greater pressures. Consequently, the pressure
waves nearer the piston tend to overtake those further from the piston.
For example, referring to t = 3, the wave in mass "a" travels faster than
the wave in mass "b," because "a" is moving faster to the right and
also because the velocity of sound in "a" is greater than the velocity
of sound in "b."
The net result of this process is that the wave profile becomes steeper
and steeper, until, a t t = 6, the pressure gradient becomes infinite.
Thus, a t t = 6, a small compression shock has been formed, a shock
which grows in strength as the process continues.
A similar analysis applied to a leftward motion of the piston, i.e., for
an expansion wave, indicates that expansion waves become less steep
with the passage of time. Thus, because of the nonlinearity of finite
wave motion, compression waves become steeper and ultimately form
a discontinuity, while expansion waves spread out and thus are unable
to support a discontinuity.
These results are summarized in Fig. 5.9, showing successive appear-
ances of a wave form of constant total pressure change, corresponding
(a) (b)
FIG.5.9. Development of wave form of constant strength.
(a) Compression wave.
(b) Expansion wave.
to the case where the piston accelerates to a constant speed and then
moves a t constant velocity; this is analogous to the case of a projectile
starting from rest and accelerating to a constant final speed. The wave
form a t a giveri instant is defined as the curve of some fluid property
(say p, p, or V ) p~dttedagainst distance 2.
STEEPENING OF COMPRESSION WAVES. The compression wave, Fig.
5.9a, becomes ever steeper as it propagates rightward, until, between
t = 4 and t = 5, i t tends to become infinitely steep a t some point. The
present analysis then ceases t o be valid because viscous and heat con-
duction effects are no longer negligible in the face of such extraordinary
gradients in velocity and temperature. Indeed, if the isentropic analysis
128 NORMAL SHOCK WAVES Ch. 5
were continued t o t = 5, the wave would "topple over" to the form shown
by the dashed line; this is physically absurd, as i t means that a t the
same time and location the fluid has simultaneously three diierent values
of pressure, velocity, and density. Viscous and heat conduction effects
intervene soon after t = 4 and act to produce a stationary shock wave
of unchanging form.
SPREADING OF RAREFACTION WAVES. The rarefaction wave, Fig.
5.9b, becomes ever less steep as it propagates rightward, and so no dis-
continuity effects are observed. Indeed, if a rarefaction shock were
momentarily established, the dynamic effects outlined here would cause
the shock to decay immediately into a continuous expansion wave.
T HE SURF ANALOGY. A similar phenomenon may be seen a t ocean
beaches. As an incoming gravity wave approaches the shore, the front
part of the wave (corresponding to compression) is seen to steepen, and
the rear part (corresponding to
rarefaction) is seen to spread out.
In this case, however, the geome-
try of the flow permits the for-
ward part to tumble over, as in
Fig. 5.9a, t = 5; this tumbling
over produces the well-known
ocean "breakers. "
Analysis of Wave Develop-
ment. The foregoing descriptive
arguments may be put into
simple analytic form.
Consider a rightward-moving
plane compression wave, the in-
stantaneous form of which is
shown in Fig. 5.10 by plots of
pressure and velocity versus dis-
FIG. 5.10. Illustrates analysis on de-
velopment of wave form. tance. At two adjacent points
the fluid properties differ i n
magnitude by dV, dp, dp, dc, etc. The respective parts of tlw
wave passing through these same points differ in wave speed by t,hc
amount dV,. Since a finite wave may be thought of as a succession of
infinitesimal pressure pulses, each element of the wave may be analyzed
as in Art. 3.2. As long as the velocity and temperature gradients are
moderate, experience confirms the assumption that longitudinal viscous
and heat conduction effects are negligible. Consequently each ele-
mentary part of the wave travels a t the local speed of sound with respect
to the fluid in which i t is propagating. The propagation velocity of a
Art. 5.6 FORMATION OF SHOCK WAVES 129
V ' l/P
- (dZp/dv2) < 0 (5.38b)
F I ~5.11.
. Criteria for steepening or The shapes of the isentropes
flattening of compression and rarefaction corresponding to these two cri-
waves in terms of curvature of p-u isen-
trope. teria are illustrated in Fig. 5.11.
Since k is positive for all known gases, it follows from Eq. 5.38a that
in a perfect gas compression waves steepen and rarefaction waves flatten.
LIQUIDS. The isentropic relation for liquids may usually be repre-
sented a t moderate pressures by
Comparing this with Eq. 5.38a, we conclude that, in liquids obeying the
law of constant p, compression waves steepen and expansion waves
flatten.
OTHER FLUIDS. NO example has yet been adduced of a real fluid obey-
ing Eq. 5.3813. However, there seems to be no basic principle forbidding
the existence of such a fluid. Possibly certain fluids in the neighborhood
of the critical pressure and temperature can exhibit stationary rare-
faction shocks.
1%-
in the equation of motion must all
-
6 --T---vx be of comparable magnitude.
Physical considerations dictate
that the stationary shock have a
form like that of Fig. 5.12. The
v ------L-- m velocity curve approaches the end
values V z and V , asymptotically,
Fro. 5.12. Definition of shock thickness. and so the thickness is actually
infinite. However, virtually the
entire change in velocity occurs in a very short distance. It is con-
venient, therefore, to define a characteristic shock thickness 6 as shown
Now, since the longitudinal viscous stress must be of the same order
of magnitude as the inertia stress,
which says that the Reynolds Number of the shock, based on the thick-
ness 6 and on the fluid properties a t T* is of the order of magnitude of
unity.
By introducing certain further relations from the kinetic theory of
gases, it may be shown that, in respect to order of magnitude,
where I is the mean free molecular path. Thus it follows that the shock
thickness is of the order of the mean free path. Thermodynamic equilib-
Art. 5.7 TH1,CKNESS OF SHOCK WAVES 133
rium does not prevail within a shock, therefore, and the analysis of a
shock from continuum considerations a t best gives approximate results
for the shock structure. Kinetic theory or even quantum mechanics
may be necessary for an adequate analysis.
Shock Thickness in Perrect Gas. The structure of the shock wave
may be investigated more fully by solving the exact Navier-Stokes
equation and the exact energy equation for such a dissipative, one-
dimensional region. The results of such studies are more accurate than
the order-of-magnitude result reached above but are nonetheless subject
to the same weakness of being based on the assumption of a continuum.
A formula showing a shock thickness Reynolds Number as a function
of Mach Number, Prandtl Number, specific-heat ratio, and viscosity-
temperature variation has been derived by Shapiro and Kline.") It
has the form
where
4 2k
D=-+--- k + 1 M , * ~+ 1
3 Pr* 2Pr* M,*
of the form p -
and where n is the exponent in a viscosity-temperature relationship
Tn (from kinetic theory, n = for perfect gases; for
air, experimental values yield n = 0.768 a t ordinary temperatures).
The symbol Pr* denotes the value of the Prandtl Number a t the tem-
perature T*. When D is positive, the plus sign in Eq. 5.39a is used;
when D is negative, the minus sign.
VERY W EAK SHOCKS. When M,* approaches unity, Eq. 5.39a may
be approximated by
V,
2 3 0 11200
V
(Ft/Sec)
- -EjNorrnol Shock
M,=I.I, T,=59F, pX*14.7psia
P/P, 1.3
1.2
P 'P,
Norrnol
Shock
/
A\@
8
'
1100 T/T, T/T,
T ~ f1 .l1 ~ #'
VY
(1054r'--q -3 -2 - 1 0 1 2 3 4 1.0
X r lo5, Inches S
+
(a) (b)
FIG.5.13. Variation of fluid properties in a normal shock.
FIG. 5.16. Schlieren photographs q f compression shock in a duct, with flow from
left to right (M.I.T. Gas Turbine Laboratory).
(a) Moderately thick boundary layer. The light, diamond-shaped regions rep-
resent regions of acceleration.
(b) Very thick boundary layer.
no~zles,the normal shock extends over most of the stream and a thick-
ening of the boundary layer occurs (Fig. 5.15a). When the shock is
strong and the boundary layer thick, as in long ducts fed by supersonic
nozzles, back flow and separation occur (Fig. 5.15b). The main stream
136 NORMAL SHOCK WAWS Ch.5
separates from the walls, alternately passes through a series of accelera-
tions and shocks, and, finally, after reaching subsonic speeds, diverges
and fills the passage again.
(a) (bj
FIG.5.17. Schlieren photographs showing the effect of boundary-layer removal on
the compression shock in a duct, with flow from left to right (after A. Weise).
(a) Boundary layer present.
(b) Boundary layer removed through suction slits.
shock has passed travels leftward with the speed Vz- V,, and, since
this is less than the shock speed V,, a particle of high-pressure gas falls
farther and farther behind the shock front.
Let us define the Mach Number of the moving shock wave as the ratio
of the speed of the wave to the speed of sound in the stationary gas.
This ratio is identical with the value of M, for the stationary discon-
tinuity. Now, since for a wave of finite strength M, is greater than unity,
Stationory Shock Moving Shock
Stationory Gas
"Y '"x
(a) (b)
FIG.5.19. Transformation of coordinate system.
(a) Gas flowing through stationary shock.
(b) Shock moving into stationary gas.
it follows that such waves travel faster than the speed of sound in the
undisturbed medium. Only in the case of a wave of infinitesimal
strength does the wave propagate with the speed of sound.
Transformation Formulas. The equations of Art. 5.3 are, of course,
valid only in terms of Fig. 5.19a1i.e., only for quantities measured by an
observer who travels with the discontinuity. In order to obtain equa-
tions which are applicable to quantities seen by an observer a t rest with
respect to the low-pressure gas, all quantities containing a velocity term
must be modified in accord with the change in coordinate system.
Suppose we signify with primes those quantities measured by an
observer who is a t rest with respect to the gas toward which the dis-
continuity moves. Then we may write
Tz' -
- Tz; T,' = T, (5.41a, 5.41b)
M,' = 0 (5.42a, 5.4213)
Art. 5.10 CONVERGING-DIVERGING NOZZLE 139
-
Po ---
LOCUS of
*
P * / P ~
States Downstream
of Normal Shock
(61
(71
(a)
IY
F I ~5.21.
. Performance of converging-diverging nozzle with various ratios of back
pressure to supply pressure.
(a) Curves of pressure versus distance along nozzle axis.
(b) Exit-plane pressure versus back pressure.
(c) Throat pressure versus back pressure.
(d) Mass flow parameter versus ratio of back pressure to supply pressure.
lowered, the shock moves down the nozzle until, at condition 4, it appears
Art. 5.10 CONVERGING-DIVERGING NOZZLE 141
in the exit plane of the nozzle. I n regime 11, as in regime I, the exit-
plane pressure p~ is virtually identical with the back pressure p ~ .On
the other hand, the flow rate in regime I1 is constant and is unaffected
by the back pressure. This is in accord with the fact that throughout
regime I1 all stream properties a t the throat section are constant.
In regime 111, as for condition 5, the flow within the entire nozzle is
supersonic, and the pressure in the exit plane 1s lower than the back
pressure. The compression which subsequently occurs outside the nozzl
involves oblique shock waves which cannot be treated on one-dimensional
grounds.
Condition 6 is termed the design condition for the nozzle under super-
sonic conditions, since the exit-plane pressure 1s then identical with the
back pressure. A reduction in the back pressure below that corre-
sponding to condition 6 has no effect whatsoever on the flow pattern
within the nozzle. In regime IV the expansion from the exit-plane
pressure to the back pressure occurs outside the nozzle in the form of
oblique expansion waves which also cannot be studied by a one-dimen-
sional analysis.
In both regimes I11 and IV the flow pattern within the nozzle is inde-
pendent of back pressure, and corresponds to the flow pattern for the
design condition. Adjustments to the back pressure are made outside
the nozzle.
For subsonic flow, there are an infinite number of possible pressure-
distance curves. For the supersonic region of flow, however, the pres-
sure-distance curve is unique. To put i t differently, in subsonic flow
the pressure ratio does not depend solely on the area ratio; in supersonic
flow the pressure ratio does depend solely on the area ratio.
Only over a narrow range of back-pressure ratios, namely, the range
covered by regime I, does the flow rate depend on the back pressure.
For regimes 11,111,and IV, the flow rate is independent of back pressure,
and, since M = 1 a t the throat, may be computed from Eq. 4.17.
0-05~ 1 2 3 4 5 6
Distance from Throol, In
A second and more serious difficulty arises. Most flow systems start
from rest and accelerate to the operating velocity. During the starting
period normal shock waves pass through the system. Across a normal
shock there is no change in flow rate or in stagnation temperature.
According to Eq. 4.17, therefore, the product Az*poz must equal the
product A,*po,. As the stagnation pressure is reduced by a normal
shock, it follows that there is a corresponding increase in the minimum
area through which the flow can be made to pass. The design problems
raised by this aspect of the process will be discussed for two applications,
namely, wind tunnel diffusers and propulsion engine diffusers.
Supersonic Wind Tunnel Diffusers. I n order to focus attention on
the essential features of the problem, it will be assumed that the flow is
isentropic, except for entropy in-
creases which occur across normal
shocks. We shall assume also that
the flow is quasi-steady during the
period of starting the flow.
STARTING CONDITIONS. Consider
the supersonic wind tunnel system
of Fig. 5.25, comprising a converg-
ing-diverging nozzle, a test section,
a diffuser, and an exhauster. We
will suppose that the inlet stagna-
tion pressure and temperature are
constant. When the exhauster is
FIG.5.25. Starting of supersonic wind started up and thereby produces a
tunnel diffuser.
reduction in pressure a t the nozzle
(a) Most unfavorable starting con- exit, a shock moves down the
dition.
(b) Best operating condition. diverging portion of the nozzle
according to the discussion of Art.
5.10. During this period, the nozzle throat is passing the maximum
possible flow, and, since A*p, is constant across the shock, it follows
that the flow can pass through the diffuser throat only if the latter is
larger than the nozzle throat. The worst condition in this respect
occurs when the shock is moving through the test section (Fig. 5.25a)
inasmuch as the shock then occurs at the maximum possible Mach
Number and, consequently, produces the largest loss in stagnation pres-
sure. The minimum area of the diffuser throat is therefore given by
section Mach Number. The limiting contraction ratio A,/A,* for the
diffuser is accordingly given by
With the limiting diffuser throat area corresponding to Eq. 5.48, the
diffuser is barely able to "swallow" the flow during starting, and the flow
a t the diffuser throat is exactly sonic when the shock is in the test
section. If the diffuser throat is slightly smaller than that required by
Eq. 5.48, a normal shock will stand in the diverging portion of the nozzle,
and subsonic flow will exist in the test section. Should the diffuser
throat be considerably smaller than that given by Eq. 5.48 the flow will
be subsonic throughout the entire system, except possibly downstream
of the diffuser throat.
The worst starting shock is illustrated in the T-s diagram of Fig. 5.27.
I
In interpreting this diagram it is well to remember that all states on the
!
I
same Fanno line have a common stagnation temperature and flow per
I unit area. A change in cross-sectional area has associated with it a
shift from one Fanno line to another. It is clear from this diagram that
I during the "starting" condition there is a large loss in stagnation pressure
and a consequent increase in the area required to pass the flow. The
path of states during the most unfavorable starting condition (i.e.,
while the shock is in the test section) is from Ox to x to y to *y to Oy.
146 NORMAL SHOCK WAVES Ch. 4
shock stands ahead of the inlet when the latter travels a t supersonic
speeds, there is no possibility of the air flow being diverted before reach-
ing the inlet, and all the free-stream flow corresponding to the cross-
sectional area Al enters the engine (Fig. 5.29b). Should the inlet not
be able to pass this amount of flow, a detached shock stands ahead of
the inlet (Fig. 5 . 2 9 ~as
) ~then the flow behind the shock is subsonic and
may spill over the inlet lip of the diffuser. Fig. 5.30 is a schlieren photo-
graph illustrating this phenomenon.
The lower curve of Fig. 5.31 shows the maximum contraction possible
when there is no shock whatsoever, and comes directly from the isen-
tropic relations. The upper curve shows the maximum contraction
possible for the special case where a detached normal shock hangs on the
inlet lip, thus forcing all the air through the engine but with a reduced
stagnation pressure. The ordinates of these curves are the reciprocals
of the corresponding ordinates of Fig. 5.26.
for O p e r a t i o n ( N o S h o c k )
I Mm
- Mad
___)
MmiMm ,
yM =I
(Shock-Free)
(c) (dl
FIG.5.33. Overspeed starting of fixed-geometry supersonic inlet, designed for free-
stream Mach Number Ma, and having contraction ratio (A2/A&.
in the acceleration period of an inlet designed for the speed M,, and
having in Fig. 5.31 a contraction ratio (A2/AI),. By overspeeding the
engine to the speed Mmd (Fig. 5.31), the shock is swallowed, and then the
engine may be decelerated at most to M,, before the shock is dissorged.
VARIABLEGEOMETRY INLET. The operating shock may also be elim-
inated through the use of a variable-geometry diffuser. Suppose that
such a diiuser is operating a t the speed M,, of Fig. 5.31. If the throat
area is too small, corresponding perhaps to point c of Fig. 5.31, a de-
tached shock stands ahead of the inlet (Fig. 5.34a). By increasing the
throat area, the shock is brought closer to the inlet. When point a of
Art. 5.11 ONEDIMENSIONAL SUPERSONIC DIFFUSERS 151
I
/'
4
-/
,
,
. ..
,. .'..:::!$:::::!;;
: ~,: ., ........ . .. . . .. . .
.!.;.$:!:.t.:*
..'.
( A t /A lo (AJA,)c
(a) (b) (c)
FIG.5.34. Starting of variable-geometry supersonic inlet, designed for free-stream
Mach Number M,.
Fig. 5.31 is reached, the shock reaches the inlet lip and may be swallowed
(Fig. 5.34b). Subsequently, the throat area may be reduced without
disgorging the shock until point c of Fig. 5.31 is reached, and, if the
exhaust pressure is properly ad-
justed, the diffuser is then free of
shocks (Fig. 5.34~).
Diffuser Efficiency. The most
common definition of diffuser effi-
ciency is parallel to the definition
employed for compressor efficiency.
Referring to Fig. 5.35, and assum-
ing that the velocity leaving the
diffuser is negligible, we define
VD =
7'3 - T1 -
T2 - Tl
T1 - - 1
(2
V12/%
)
and, since k-l
-
T3/7'1 = (p2/pi)
and
we get k- 1
152 NORMAL SHOCK WAVES Ch.5
The efficiency is uniquely related t o the stagnation pressure ratio
for a given inlet Mach Number. To demonstrate this, we write
-P2 = Po1
- .P- z Po1
= -Po2. -
Pl Pl Pol PI Pol
Employing the isentropic relation for p o l / p l , and substituting into Eq.
5.50, we get after rearrangement,
k-1
r l ~=
(I+? M12) ?):( 1
(5.51)
k-1
-M12
2
Corresponding to the two curves of Fig. 5.28, there are plotted in
Fig. 5.36, with the aid of Eq. 5.51, the maximum efficiencies for fixed-
geometry diffusers and normal shock diffusers.
Fig. 5.39 shows this relation for a gas with k = 1.4. T h e ratio pov/pz
is also tabulated in Table B.3.
Subsonic Pitot Tube. For subsonic flow, there is no shock (unless
M , is close ta unity), and Eq. 5.52 is not applicable. Instead, the cus-
tomary isentropic relation, Eq. 4.14b, is used for relating the Mach
Number, the static pressure, and the isentropic stagnation pressure.
Measurement of Static Pressure. I n using Eq. 5.52, t h e static pres-
sure p, must be measured upstream of t h e shock wave. If the stream
is in a wind tunnel, this might be done with a static pressure t a p in the
wall of the tunnel.
Pitot tubes often have a static t a p built into the side wall of the probe.
I n supersonic flow this t a p does not give a n easily interpretable measure-
ment. However, experimental data indicate t h a t if the static t a p is
placed some ten tube diameters aft of the nose, the static pressure at the
t a p is a close approximation t o the static pressure upstream of t h e shock.
PROBLEMS
5.1. Air flows steadily through a pipe of constant cross-sectional area and
at a certain section has properties p, = 10 psia, T , = 124OoF,and V, = 3000
ft/sec.
(a) Plot the Fanno and Rayleigh lines corresponding to point x on the p,
T-v, and T-s diagrams. Find from these diagrams the pressure, temperature,
and velocity downstream of a normal shock occurring at condition x.
(b) Compare your results with the corresponding results found from the
shock tables.
5.2. Show that for a very weak shock in air,
Art. 5.12 PROBLEMS 165
6.3. Derive simple approximate formulas for very weak shocks, expressing
(1 - (P, - P=)/P=, (Tv - l1J/Tz, (V, - Vz)/Vz, and (PO*- P ~ / P O aU
*
in terms of (M, - 1).
5.4. (a) Find the limiting forms of the normal shock equations for a perfect
gas with k = 1. (b) Find simple asymptotic forms of the normal shock equa-
tions for a perfect gas when the Mach Number approaching the shock is verg
large compared with unity.
6.5. (a) Show that the equation of the Fanno line for a perfect gas is
where s, and T , denote respectively the entropy and temperature at the point
of maximum entropy on the Fanno line.
(b) Show that the equation of the Rayleigh line for a perfect gas is
where sa and Tt,are respectively the entropy and temperature a t the point of
maximum entropy on the Rayleigh line. Demonstrate that there is a point of
maximum temperature on the Rayleigh line and find a relation connecting the
entropy and temperature at this point with sb and Tb.
5.6. Consider a normal shock in a fluid (not a perfect gas) whose density
depends only on the pressure according to the relation
dp
P-=8
d~
where B is a ~ositiveconstant.
(a) show that the inlet and exit Mach Numbers are related by
5.11. Suppose that a blast wave, which might have been initiated by an
atomic bomb explosion, is traveling through air a t standard atmospheric condi-
tions with a speed of 200,000 ft/sec.
Estimate the changes in pressure (atm), temperature ( O F ) , stagnation pressure
(atm), stagnation temperature (OF), and velocity (ft/sec), ~roducedby the wave
with respect to an observer who is stationary with respect to the undisturbed air.
5.12. A particular convergingdiverging nozzle is designed for a Mach
Number of 2 on the basis of reversible, adiabatic flow. When the ratio of
exhaust-region pressure, p,, to supply-region pressure pa, is raised considerably
above the design value, a normal compression shock appears in the nozzle. The
pressures before and after the shock will be denoted by p, and p,, respectively.
Assuming reversible, adiabatic flow both upstream and downstream of the
shock, plot p,/po and pulpoagainst p,/po for the range in which the shock lies
between the throat and the exit plane. The shock may be considered as being
infinitesimal in thickness, and the fluid may be taken as a perfect gas with
k = 1.4.
Compare the curve of P,/P, with a curve representing the isentropic pressure
ratio a t the same location in the nozzle for the limiting case of subsonic operation,
that is, for the case where the Mach Number is unity a t the throat and the flow
in the diverging section is subsonic.
5.13. A rocket nozzle which is intended to operate a t an altitude of 100,000
ft has an area ratio of 5.16. At launching, however, the rocket exhausts to nor-
mal atmospheric pressure, 14.7 psia. The rocket gases are generated a t 400 psia
and have k = 1.3. Assuming frictionless, adiabatic flow in the nozzle, find the
ratio of the rocket thrust a t launching to the thrust a t 100,000 ft.
6.14. A fixed-geometry, convergent-divergent wind tunnel diffuser is to be
designed for Mach Number 2. Assuming no friction, compare the maximum
possible efficiencyand the minimum possible per cent loss in stagnation pressure
during operation for the following cases:
(a) The best possible design is employed.
(b) The design is conservative, with a throat area 5 per cent larger than
that required for starting, and with the shock located during operation
a t an area 5 per cent greater than the throat area.
(c) The converging portion is eliminated, and the process comprises a
normal shock followed by reversible subsonic compression.
5.15. Consider a fixed-geometry, converging-diverging wind tunnel diffuser
in which the minimum area is 15 per cent smaller than the entrance area, and
in which the efficiency for subsonic deceleration to zero velocity is 90 per cent.
Frictional effects upstream of the shock are negligible.
(a) Plot the maximum possible efficiency and the minimum possible per cent
loss in stagnation pressure versus the approach Mach Number. What is the
minimum value of the latter for starting purposes?
(b) Assume that during operation, because of conservative design, the shock
stands downstream of the throat a t a point where the area is 10 per cent larger
than the throat area. Plot the efficiency and the per cent loss in stagnation
pressure versus the approach Mach Number.
Ch. 5 PROBLEMS IS7
5.16. A ram jet aircraft is to fly a t 40,000 ft altitude with a speed of 2000
mah.
(a) Design the best fixed-geometry, convergent-divergent diffuser for this
aircraft, and compute for it the maximum efficiency and the least per cent loss
in stagnation pressure.
(b) Suppose that it were possible to overspeed the aircraft to 2400 mph.
Design the best convergent-divergent diffuser which could then be used, and find
for it the maximum efficiency and the least per cent loss in stagnation pressure
a t the operating speed of 2000 mph.
sections.
(b) Indicate on the sketch the pressure (psia) and temperature ( O F ) a t the test
section, a t the entrance to the compressor, and a t the exit of the compressor.
(c) specify the diameter of the nozzle throat.
(d) Specify, for the compressor, the pressure ratio, the volume rate of flow at
inlet (cfm), and the horsepower required for operation.
5.18. A certain pitot tube is used for measuring the Mach Number M, of a
supersonic air stream, and the pressures shown in the sketch are recorded. It
is known that for the particular tube employed, the
pressure coefficient (referred to free-stream condi-
tions) a t the static holes is 0.1. Estimate the free-
stream Mach Number.
5.19. When an impact tube with a blunt nose is
placed in a supersonic stream, a curved shock front
4~~
shock Holes
f p - 2 0 psi0
PROB.5.18.
stands a t some distance in front of the tube. An
impact-tube traverse in a wind tunnel gives values of 2.34 psia and 10.02 psia
for the static pressure upstream of the shock and the pressure a t the mouth of
the tube, respectively. Estimate the Mach Number of the tunnel.
158 NORMAL SHOCK WAVES Ch. 5
5.20. Explain why entropy decreases occur as a fluid particle nears the end
of its passage through a shock wave.
5.21. (a) Consider a rightward moving wave in a perfect gas. Show that,
one follows a part of the wave corresponding to a particular fluid velocity
rts
and pressure, the time rate of change of wave steepness is given by
(b) Show that the time required for a given part of the wave to acquire infinite
slope is given by
2
~IOY
5.24.A jet pump (see sketch) uses a high-pressure stream of air to pump a
second stream of atmospheric air to a pressure of 18.25 psia. The mixing tube
of this particular pump has constant cross-sectional area. As a first approxima-
tion, all irreversibilities except those associated with viscous mixing are to be
ignored.
Calculate the mass rates of flow of primary air and secondary air.
Chapter 6
FLOW IN CONSTANT-AREA DUCTS
WITH FRICTION
NOMENCLATURE
p
T,,,
( )*
density
wall shearing stress
signifies state at which M = 1
I ( )o
signifies state at which M
l/dkin isothermal flow
signifies stagnation state
=
where p is the density a t the section where V and h are measured, and
G (the mass velocity) has a constant value for all sections of the duct.
Combining Eqs. 6.1 and 6.2, we get the equation of the Fanno line
in terms of the enthalpy and density:
Since ho and G are constants for a given flow, Eq. 6.3 defines a relation
between the local enthalpy and the local density. This relation is
shown graphically'in Fig. 6.1 for a single value of ho and for several
values of G. All possible states of the fluid for a given adiabatic, con-
stant-area flow lie on one of these lines.
THE FANNO LINE. For a pure substance, s = s(h,~),i.e., the entropy
is determined by the enthalpy and density., The curves of Fig. 6.1
may thus be transferred to the enthalpy-entropy diagram, giving the
Art. 6.1 INTRODUCTORY REMARKS 161
familiar Fanno curves of Fig. 6.2. For all substances thus far investi-
gated, the Fanno curves have the general shape shown in Fig. 6.2.
It was shown in Art. 5.2 that the upper branch of each Fanno curve
corresponds to subsonic flow, that the lower branch corresponds to super-
sonic flow, and that the Mach Number is unity a t the point of maxi-
mum entropy on each Fanno curve.
SECOND-LAW LIMITATIONS. Since the flow is adiabatic, the Second
Law of Thermodynamics tells us that the entropy may increase but
'\ \i
FIG.6.1. Fanno lines on h-v diagram. FIG.6.2. Fanno lines on h-s diagram.
Solid lines have the same stagnation en- The three curves shown have the same
thalpy but different flows per unit area. stagnation enthalpy but different flows
Dashed lines are isentropes. per unit area.
may not decrease. Thus, the path of states along any one of the Fanno
curves must be toward the right.
Consequently, if the flow a t some point in the duct is subsonic (point a
of Fig. 6.2), the effects of friction will be to increase the velocity and
Mach Number and to decrease the enthalpy and pressure of the stream.
If, on the other hand, the flow is initially supersonic (point b of Fig. 6.2),
the effects of friction will be to decrease the velocity and Mach Number
and to increase the enthalpy and pressure of the stream. A subsonic
flow may therefore never become supersonic, and a supersonic flow
may never become subsonic, unless a discontinuity is present. Thus
we observe that, as in the case of isentropic flow, the qualitative char-
acter of the flow is markedly influenced by whether the flow is subsonic
or supersonic.
The limiting pressure, beyond which the entropy would suffer a d e
crease, occurs a t Mach Number unity and is denoted by p*. The asterisk
here denotes the state where M = 1 for the particular process under
consideration, namely, adiabatic flow a t constant area. Thus, referring
162 FLOW IN CONSTANT-AREA DUCTS WITH FRICTION Ch.6
to state a as an example, the value of pa* will be diierent for an isentropic
flow as compared with the value for an adiabatic, constant-area flow.
Fig. 8.2 indicates that the isentropic stagnation pressure is reduced
as a result of friction, irrespective of whether the flow is subsonic or
supersonic.
CHOKING DUE TO FRICTION. Consider a situation in which the stag-
nation enthalpy, flow per unit area, and length of duct are such that
Mach Number unity is reached a t the end of the duct. If the duct
length is increased, i t is evident from the foregoing considerations that
some sort of adjustment in the flow is necessary. When the flow is
subsonic, this adjustment is in the form of a reduction in the flow rate,
that is, the flow is choked. When the flow is supersonic, the adjustment
at first involves the appearance of shock waves, and, for sufficiently
large increases in duct length, involves ultimately a choking of the flow.
This will be discussed more fully in Arts. 6.2 and 6.3.
Conrrol S u r f a c e
p = pRT
we take logarithmic differentials, and thus obtain
Dividing through by c,T, and using the definition of the Mach Number,
this becomes
dT k - 1 dV2
T
-+- 2
~2 --
V2
-0
We now introduce the latter two expressions and the continuity equa-
tion into the momentum equation, and thus obtain
Subsonic Supersonic
Pressure, p decreases increases
Mach Number, M increases decreases
Velocity, V increases decreases
Temperature, T decreases increases
Density, p decreases increases
Stagnation Pressure, po decreases decreases
Impulse function, F decreases decreases
We see that the Mach Number always tends toward unity. Con-
tinuous transitions either from subsonic to supersonic flow, or from
supersonic to subsonic flow, are consequently impossible. For given
conditions a t an initial section of the duct, therefore, the maximum
possible duct length which can be employed without altering these given
initial conditions and without introducing discontinuities is that length
for which the exit Mach Number is exactly unity.
It is a t first surprising to see that friction has the net effect of accelerat-
ing a subsonic stream, and perhaps even astonishing t o observe that
friction causes a rise in pressure a t supersonic speeds.
Working Formulas. Our next step is to integrate the previously given
differential equations in order t o obtain formulas suitable for practical
Art. 6.2 ADIABATIC FLOW OF A PERFECT GAS 167
where the limits of integration are taken as (i) the section where the
Mach Number is M, and where x is arbitrarily set equal to zero, and
(ii) the section where the Mach Number is unity, and x is the maximum
possible length of duct, Lmax.
Carrying out the integration, we obtain
f =-i
1
Lmax
Lm,
jdz
Eq. 6.20 gives the maximum value of 4fL/D corresponding to any initial
Mach Number M.
Since 47LmaX/Dis a function only of M, the length of duct L required
for the flow to pass from a given initial Mach Number M1 to a given
final Mach Number Mz is found from the expression
nlustrative Example.
P R o B ~ M With
. an experimental rig comprising a converging-diverging n o s
zle attached to a smooth, round tube, the following data were obtained with the
aim of measuring friction coefficients for the supersonic flow of air:
Pressure upstream of nozzle = 516 cm Hg abs
Temperature upstream of nozzle = 107.3"F
FLOW IN CONSTANT-AREA DUCTS WITH FRICTION Ch. 6
(4fLmar/D)1= 0.4371
Now, since
-
P2 = ( P / P * ) z
PI (P/P*)I
we obtain
37.1
( P / P * )=~ 18.25(0.2878) = 0.5850
(4fLm,/D)? = 0.1512
Therefore,
AL
4f - = 0.4371 - 0.1512 = 0.2859
D
and, since
AL/D = 29.60 - 1.75 = 27.85
the average friction coefficient is
shock. In Fig. 6.6 the curves labeled B and C represent the two extreme
locations: B referring to a normal shock a t the exit of the duct, and C
referring to a normal shock a t the inlet to the duct.
It is clear from Fig. 6.6 that for any initial Mach Number there is a
maximum value of 4jL/D for which a solution is possible, and that
M, = 1 a t this condition. From a different point of view, for a given
value of 4jL/D there is in subsonic flow a maximum initial Mach Number
for which a solution is possible, and in supersonic flow a minimum initial
Mach Number for which a solution may be found. It is pertinent to
ask, therefore, what happens when the duct length initially has its
maximum value for a given MI and the duct length is subsequently
increased. In answering this question we shall suppose that the back
pressure to which the duct exhausts is maintained as low as necessary,
so that any diminutions in the flow are due exclusively to limiting effects
produced by friction.
SUBSONIC FLOW. An increase in the value of 4fL/D over its maximum
value will act to decrease MI until a steady-state solution again becomes
possible with M2 = 1. This results in a reduction in the flow rate, i.e.,
the flow is "choked" by friction.
SUPERSONIC FLOW. An increase in the value of 4jL/D over its maxi-
-
mum value a t first produces a shock in the duct which moves upstream
as the duct length is increased. In
other words, for a fixed MI and a
variable 4jL/D, the value of Ma
remains a t unity, but the flow
pattern in the duct moves from a
C
condition corresponding to curve
,y
&
rt-..-:-=
"To I
L I I
0 Distance Along Duct
Fro. 6.9. Flow passed by long adiabatic duct fed by isentropic converging nozzle
(k = 1.4).
(a) Dimensionless flow parameter versus ratio of exhaust pressure to supply
pressure.
(b) Ratio of choking flow to choking flow through nozzle alone, as function of
.4fL/D.
Since the nozzle is without shocks, the pressure and Mach Number a t
section I are fixed by the area ratio of the nozzle. To the Mach Number
MI corresponds a particular value of 4jLm,,/D according to Eq. 6.20.
We therefore divide the types of flow into two main classes, depending
CLASS (2). The duct length is greater than the maximum possible
length, so that p p > p* and L > L,,, (Fig. 6.12).
The value of p* is a function only of MI and PI. The value of p ~ ,on
,
the other hand, depends on MI, p ~and, 4jL/D.
Art. 6.3 PERFORMANCE OF LONG DUCTS 177
p~ > p*, the flow a t the duct exit must be subsonic, and, therefore, the
exit pressure p~ must be identical with the back pressure p ~ .This
condition, namely, that p~ = p ~ determines
, the location of the shock
in the tube.
Increases in the back pressure cause the shock to move upstream and
ultimately to vanish in the throat of the nozzle.
CLASS (2b): p p > p*, and p~ < p* (Fig. 6.12~). As in the case of
Class (2a), the duct length is so great that i t is impossible for the flow
to be supersonic in the entire length of duct. Accordingly, a shock is
present.
The flow a t the duct exit cannot be supersonic, for the flow downstream
of the shock is subsonic and can, therefore, not become supersonic
except through area chalige or heat transfer.
Furthermore, the flow a t the duct exit cannot be subsonic, with a
pressure greater than p*, for on leaving the duct it would diverge on
account of the exit pressure being greater than the back pressure, and
the pressure increase produced by this divergence would make it impos-
sible for the stream ever to adjust itself to the back pressure.
The sole remaining possibility is that the stream leaves the duct at
Mach Number unity, or, in other words, that p~ = p*. The adjustment
of the stream to the back pressure then occurs outside the duct in the
form of oblique expansion waves. As long as the back pressure is less
than p*, therefore, there is only one location in the duct a t which the
shock may stand.
Finally, it should be mentioned that the flow patterns of Figs. 6.10
and 6.11 are idealized in that the shocks are shown as discontinuities.
Because of boundary-layer effects, the transition from supersonic to
subsonic flow is more likely to be extended over a considerable distance,
as explained in Chapter 5 .
~ P / P= ~ P / P
Similarly, Eq. 6.5 becomes
Eqs. 6.7, 6.8 and 6.9 are valid as they stand for isothermal flow.
positive, the directions of change are summarized for gases with k >1
in the table which follows.
M<I/& M> I/&
(Subsonic) (Subsonic or Supersonic)
Pressure decreases increases
Density decreases increases
Velocity increases decreases
Mach Number increases decreases
Stagnation temperature increases decreases
increases for M <
Stagnation pressure decreases
decreases for M >
It is seen that the Mach Number always tends towards I/%'% This
value therefore represents a limit for continuous isothermal flow, in the
same way that M = 1 represents a limit for continuous adiabatic flow.
When M is less than ljdi,heat is added to the stream; when M exceeds
I/&, heat is rejected from the stream.
Working Formulas. Eq. 6.33 may be rearranged t o give
from which
,~
The perfect-gas relation yields
Art. 6.4 ISOTHERMAL FLOW I N LONG DUCTS
From the formula for isentropic stagnation pressure, we get
M
FIG.6.13. Working chart for isothermal flow in ducts, k 3 1.4.
182 FLOW IN CONSTANT-AREA DUCTS WITH FRICTION Ch. 6
1 - kM12 1 - kMZ2
-
m12
-
kMZ2
+ ln-M12
~2~
L
D
4f-=
1-T;(
~ , 2- ln (E)2
(6.42)
Hence, although for given values of MI and 4fL/D there are two solu-
tions for pz/pl, only the left-hand intersection may be used. The right-
hand intersection implies negative values of f and involves a violation of
the Second Law of Thermodynamics. A tangent-type intersection
indicates choking flow, and, when there is no intersection a t all, L
exceeds Lm,.
When the per cent pressure drop is fairly small, i t is convenient to
expand the right-hand side of Eq. 6.42 in a power series of the fractional
pressure drop, (pl - p2)/pl. Carrying out this expansion, and retaining
terms up to the second power of this variable, we obtain the following
useful approximation applicable to low pressure drops:
Art. 6.4 ISOTHERMAL FLOW I N LONG DUCTS 183
use L e f t - H a n d
Intersection On
P,/P, Scale
FIG. 6.15. Friction coefficient versus pipe Reynolds number for incompressible,
fully developed flow. Roughness of pipe is measured by e, and has typical values as
follows (after Moody):
P
- be c. f t
-
Drawn tubing 0.000005
Commercial steel 0.00015
Asphalted cast iron 0.0004
Galvanized iron 0.0005
Cast iron 0.00085
Concrete 0.0014.01
Riveted steel 0.0034.03
For a fully developed velocity profile, i.e., a t distances from the duct
inlet greater than about 50 diameters, no significant effect of Mach
Number was observed. In other words, the relation between friction
coefficient and pipe Reynolds Number for subsonic flow a t any Mach
Number was found to agree with the well-known Karman-Nikuradse
formula for incompressible flow:
PROBLEMS
6.1: A stream of air flows in an insulated tube of constant area having a
cross-sectional area of 1 sq ft. At a section 1 the pressure is 10 psia, the temper-
ature is 40°F, and the mass velocity is 29.6 lb/sec ft 2 . The pressure in the space
to which the tube exhausts is so low that a "choking" condition prevails, i.e.,
alterations in the pressure of the exhaust region have no effect on the flow through
the tube.
(a) Calculate the Mach Number at section 1.
(b) Calcuhte the Mach Number, temperature, and pressure at the exit of
the tube.
(c) Calculate the total force in the axial direction which must be exerted to
hold statbnary the section of the duct between section 1 and the exit.
6.2. A perfect gas flows in an insulated pipe. Show that the pressure at'the
choking condition is given by
Art. 6.5 PROBLEMS 187
6.3. (a) Estimate the maximum flow rate of air (lbm/sec) which can flow
through the passage shown, assuming that the friction coefficient of the duct
is 0.005. (b) For what range of back pressures will
this maximum flow rate be achieved?
6.4. Consider a long, round insulated tube fed
with air by a frictionless nozzle. The critical pres-
sure ratio of this system is defined as the ratio of PROB. 6.3.
the minimum possible pressure a t the duct exit to
the nozzle supply pressure. Plot this critical pressure ratio versus the duct
4fL/D for values of the latter from 0 to 10.
6.5. Air enters an insulated tube of 1-inch diameter through a converging-
diverging nozzle with a throat diameter of 0.400 inches. The pressure and
temperature in the low-velocity region at the nozzle entrance are 100 psia and
140°F, respectively. The flow through the nozzle may be assumed frictionless,
and the average friction coefficientin the duct may be taken as 0.0025.
(a) Plot the pressure, temperature, Mach Number, and isentropic stagnation
pressure as a function of distance from the tube inlet, assuming shockless flow
to the exit. Determine the maximum length of duct for these conditions.
(b) Suppose that the tube is 40 inches long. Plot, versus the back pressure
in the space to which the duct discharges, the following quantities: mass rate
of flow, pressure and Mach Number at the duct exit, pressure and Mach Number
at the duct inlet, and location of shock.
(c) Repeat part (b) for a tube 90 inches long.
6.6. An isentropic nozzle having an area ratio of 2.00 discharges air into
an insulated pipe of length L and diameter D. The nozzle is supplied a t 100
psia and 70°F, and the duct discharges into a space where the pressure is 40
psia. Calculate the 4jL/D of the pipe and the mass flow per unit area in the
pipe.{&m/sec ft2) for the cases where a normal shock stands:
(a) In the 'nozzle throat
(b) In the nozzle exit plane
(c) In the duct exit plane
6.7. (a) Demonstrate that for incompressible flow of a perfect gas in a pipe
the fractional pressure drop is given by
(b) Derive the corresponding expression for the fractional loss in stagnation
pressure.
6.8. Natural gas (molecular weight 18, k = 1.3) is to be pumped through
a pipe of 36 inches I.D. connecting two compressor stations forty miles apart.
At the upstream station the pressure is not to exceed 90 psig and at%hedbwn-
stream station it is to be at least 10 psig. Calculate the maximum allowable
rate of flow (cubic feet per day at 70°F and 1 atm), assuming that there is
sufficient heat transfer through the pipe wall to maintain the gas a t 70°F.
188 FLOW IN CONSTANT-AREA DUCTS WITH FRICTION Ch. 6
6.9. A converging nozzle supplied with air at 100 psia and 70°F discharges
into a duct 10 inches in diameter and 1000 ft long. Assuming the flow in the
nozzle isentropic and the flow in the duct isothermal,
(a) Plot the mass rate of air flow (lb/sec) versus the exhaust pressure (psia)
in the space to which the duct discharges.
(b) What is the maximum mass rate of flow?
(c) What is the rate of heat transfer (Btu/hr) to the duct for the maximum
flow rate?
(d) What is the inlet Mach Number to the duct at the maximum flow rate?
6.10. Consider isothermal flow through a frictionless nozzle supplied with a
gas at low velocity, pressure pi, and temperature T,.
(a) Derive relations for p/pi, p/pi, V/V*t, To/Toi, po/poi, and A/A*" all in
terms of the local Mach Number M, where To and po are respectively the isen-
tropic stagnation temperature and pressure, and V*l and A*' are respectively
" and cross-sectional area at the section where w/A is a maximum.
the velocitv
(b) Show that at the minimum area the Mach Number is l/& and the
pressure ratio is p*l/p; = I/& = .-/I
(c) In what way are the results of parts (a) and (b) identical with what would
be found from the isentropic-flow relations by setting k = l ? Investigate the
reasons for the similarities and dissimilarities.
6.11. Derive a formula relating (T ds - dQ) to 4f dx/D for the isothermal
pipe flow of a perfect gas. From this show that the friction coefficient f must
always be positive.
6.12. The measurements shown on the sketch are made in a certain wind
tunnel where it is known that the boundary layers on the walls of the test
section are very thin.
In interpreting test data on the model, what would you consider the free-
stream Mach Number for the model? Defend your conclusion.
the gas has properties k = 1.4 and R = 53.3 ft lbf/lbm F. The viscosity of the
gas may be taken as 2.6(10)-5 lbm/ft sec. Since the flow in the capillary is
likely to be laminar, the friction coefficient is given by the Poisseuille formula:
f = 16/Rey
(a) Estimate the maximum rate at which gas may be sampled, in pounds per
second.
(b) How much vacuum must the pump develop, in inches Hg, in order to
realize the maximum flow rate?
(c) What is the pressure a t the inlet of the capillary a t the condition of
maximum flow?
6.14. Consider the steady, laminar seepage of a perfect gas through a long,
fine capillary tube of diameter D and length L. Because of its very great length
and small diameter, the resistance to flow is very large, and consequently the
fluid velocities are so small that changes in fluid momentum and kinetic energy
are negligible. At the same time, the pressure drop is so large that there are
great variations in fluid density.
Show that the pressure drop is given by "d'ArcyJs formula,"
NOMENCLATURE
area (R recovery factor
speed of sound s entropy per unit mass
specific heat at constant pressure T absolute temperature
pipe diameter Taw adiabatic wall temperature
impulse function T, wall temperature
mass velocity V velocity
enthalpy per unit mass w mass rate of flow
latent heat per unit mass
heat of reaction at constant X coefficient of thermal conductiv-
temperature and pressure per ity
unit mass of mixture p coefficientof viscosity
heat transfer coefficient p density
ratio of specific heats
Mach Number ( ) * signifies state at which M = 1
pressure for process of simple To-
rate of heat transfer per unit Change
wetted area ( )o signifies isentropic stagnation
heat state.
gas constant
T h e equation of continuity is
w
pV = - = G = constant
A
Combining these, we have
The Rayleigh Lie. For nxed values ot,the now per unit area and t h e
impulse function per unit area, Eq. 7.3 defines a unique relation between
the pressure and the density, a relation which we have previously called
the Rayleigh line. Since both the enthalpy and entropy are functions of
192 FLOW I N DUCTS WITH HEATING OR COOLING Ch. 7
pressure and density, it follows that Eq. 7.3 may be used for representing
the Rayleigh line on the enthalpy-encropy diagram, as in Fig. 7.1.
Writing Eq. 7.3 in differential form and introducing Eq. 7.2, we get
dp G2
-=-- - V2
3
dp p2
or,
v = M
Now, eP
represents the local velocity of sound only for a special
circumstance, namely, when the
/PO, Ray'eigh, infinitesimal variation of pressure
h3L-+---
' dition
no
withchange
density
is fulfilled
of isentropy.
such
a t that
theThis
point
there
con-
of
is
Number is unity. If the heat addition is too great, the flow will be
choked, that is, the initial Mach Number will be reduced to a magnitude
which is consistent with the specified amount of heat input.
These phenomena will be investigated in detail in the articles which
follow.
7.2. Simple-Heating Relations for a Perfect Gas
As in the preceding chapters, we shall pay special attention to the
flow of a perfect gas. The inaccuracies which are thus introduced are
in most cases more than justified by the resulting simplicity.
Governing Physical Equations and Definitions. Consider the flow
through the control surface of Fig. 7.2. We shall write the physical
i T0 2
1
P
L -----------------J
FIG.7.2. Control surface for analysis of Simple TO-Change.
Mam NUMBER. Using the definition of the Mach Number and the
formula for the sound velocity in a perfect gas, we find
ISENTROPIC
STAGNATION PRESSURE. The definition of the isentropic
stagnation pressure, Eq. 4.14b, gives
ENTROPY CHANGE. The entropy change may be found from Eq. 2.32~.
Thus,
- -
-- (7.16)
To* (1 + kM2)2
Similarly, we obtain the following formulas:
(7.17)
(7.18)
(7.19)
(7.20)
Illustrative Example.
PROBLEM. A gaseous mixture of air and fuel enters a ram-jet conlbustion
chamber with a velocity of 200 ft/sec, a t a temperature of 120°F, and a t a pres-
sure of 5 psia. The heat of reaction AH of the mixture for the particular fuel-
air ratio employed is 500 Btu per pound of mixture. It is desired to find the
stream properties a t the exit of the combustion chamber. I t will be assumed
that friction is negligible, that the cross-sectional area is constant, and that the
properties of both the reactants and the products are equivalent to air in respect
to molecular weight and specific heat.
and
MI = Vl/cl = 200/1180.3 = 0.16945
With this value of (To/To*)2,we again enter Table B.5 and find that
( P / P * ) z- 5-
1.9012 = 4.119 psia
Pz = Pl
( P / P * ) ~ 2.308
(PO/PO*)Z 1.1425
= 5.102 -= 4.688 psia
Po2 = Pol
(PO/PO*)~ 1.2435
It is also of interest to find the maximum heat of reaction for which flow with
the specified initial conditions may be maintained. Choking will occur when the
exit Mach Number is unity. Hence, denoting the choking condition by sub-
script 3,
(T0/T0*)3 1
= 583.0 -= 4550°R
= ( T O / T ~ * ) ~ 0.128
After the fuelair ratio has been increased to the value corresponding to this
maximum heat of reaction, further enrichment of the mixture will produce a
reduction in mass rate of flow and a consequent reduction in the initial Mach
Number.
Effects of Changes in To. T h e effects of changes in To on the
remaining stream properties may b e found from differential relations
of the type derived in t h e previous chapter. Such differential relations
are derived a s part of a more general analysis in the next chapter and so
are not presented here.
However, the desired information may also be found from Fig. 7.3,
a t least for the special value of k = 1.4. T h e key curve on this chart
is t h a t of To/To*,for it is directly connected with the magnitude and
Art. 7.2 SIMPLEHEATING RELATIONS FOR A PERFECT GAS 199
To I
-
To
Violates
2nd Law
---Expansion Shock
0 +Compression Shock
0 I 2
0
0 I 2'
FIG. 7.6. Schlieren photographs (flow left to right) of moisture condensation shock
in air. The moisture shock is immediately downstream of the throat; as i t is not quite
normal near the wall, there is a reflected expansion wave a t each wall. Further down-
stream of the moisture shock is an adiabatic normal shock with attendant boundary-
layer separation (M.I.T. Gas Turbine Lab.).
(a) Schlieren knife-edge vertical.
(b) Schlieren knife-edge horizontal.
the shock, W L the mass rate of flow of condensing liquid, and h,, the
latent heat per unit mass of condensing liquid.
Eq. 7.22 may be rearranged to give
FIG.7.7. Curves of M2 versus MI for values of the parameter AH/c,TI in the usual
range of hydrocarbon explosions. Dotdash curves indicate values of pz/pl ( k = 1.4).
-
tions discussed below are illustrated in Figs. 7.7 and 7.8.
Front Front
- -
(Stationary)
-4
'
VI
I
I V,
@
Vl
Unburned
( A t Rest) I
@ Burned
- -
I I -v2-v,
@ Unburned I @ Burned p~'P2 I
I Vl
Front Front
"1 Vl
a Unburned
( A t Rest)
I @
I
Burned @ Unburned
( A t Rest)
I @ Burned
FIG.7.9. Thermally-choked combustion waves (M2 = I), fork = 1.4 Solid curves
show maximum flame speed MI for slow combustion, and minimum flame speed M I
for detonation. The latter is the detonation Mach Number always observed in experi-
ments. Dashed curve shows, for detonation with M2 = 1, the stagnation pressure
ratio for an obsewer a t rest with respect to the unburned gas.
The two values of M I which may be found from this equation corre-
spond to the maximum flame-front speed for slow combustion and the
Art. 7.5 THE RECOVERY FACTOR 21 1
minimum speed for detonation. These speeds are plotted in Fig. 7.9 as
a function of the energy parameter AH/c,Tl.
One measure of the potentialities for damage of a steady detonation
wave is the rise in stagnation pressure experienced by an obstacle over
which the wave passes. Denoting by pbzl the stagnation pressure of the
burned gas in a coordinate system where the unburned gas is a t rest,
it may be shown that
thus tending to limit the temperature rise. After these effects have
been brought into balance everywhere in the fluid, the wall is a t the
equilibrium temperature Taw.The
temperature distribution in the
boundary layer is then as shown
-E0,. I Oistrfbutlons
by the solid curve of Fig. 7.10.
0
I I Cold Woll Corresponding curves for a hot wall
a
*" .-. \
I Insulated W a l l
and a cold wall are also shown; it
0
-L"
c
XO'
.. Wall should be noted that whether heat
0 . .*
,' '. flows into or out of the wall depends
T-o-r
Temperature
on the direction of the temperature
FIG. 7.10. Illustrates definition of
gradient in the gas a t the gassolid
adiabatic wall temperature. boundary.
The Recovery Factor. The magnitude of Tawrelative to the one-
dimensional mean values of T and To is usually expressed by the so-
called recovery factor &, which is defined by
T
-aw- Taw
- -
rn
1 1
-T
so that
Moreover, i t is shown in Volume 11, Chapters 26,27, and 28 (cf. Fig. 28.4)
t h a t the value of X defined in terms of (Taw- T,) is the only one lead-
ing t o an experimental correlation which is independent of the rate of
heat transfer. Accordingly, we define X b y
where q is the rate of heat flow per unit wetted area. Substituting for
Tawwith the aid of Eq. 7.30, the heat transfer equation becomes
PROBLEMS
7 (a) Fuel is to be burned in a stream of air moving with a Mach Number
of 2.0, a temperature of - 160°F, and a pressure of 10 psia in a duct of constant
area. Estimate the maximum possible rise in stagnation temperature, and the
corresponding final values of the temperature, pressure, and stagnation pressure.
(b) If the supersonic stream of part (a) were to pass through a normal shock
before fuel was burned, estimate the maximum possible rise in the temperature,
214 FLOW I N DUCTS WITH HEATING OR COOLING Ch.7
and the corresponding final values of the temperature, pressure, and stagnation
pressure.
7.2. Make log-log plots of M2 versus MI, p2/pl versus MI, T2/T1 versus Mi,
and po,/po, versus Mi, for fixed values of To2/TO1
of 1.2 (heating) and 0.8 (cool-
ing). Note that there are two branches on each plot for each value of TodToi.
Use solid lines to indicate which portion of each branch is physically possible
and dashed lines to indicate which portions involve violations of the Second
Law of Thermodynamics.
7.3. Derive the asymptotic forms of the Simple To-Change formulas for:
(a) Very low Mach Numbers
(b) Very high Mach Numbers
7.4. Show that, in Simple To-Change a t Mach Numbers less than about 0.2,
changes in density are produced almost exclusively by changes in temperature.
Based on this observation, and considering two sections 1 and 2, demonstrate
that at very low Mach Numbers,
7.5. Fuel is injected and burned in air moving down a straight pipe of uniform
diameter. If the constant-pressure heat of reaction of the fuel is 19,000 Btu/lb,
the stagnation temperature of the air a t inlet is 90OoR, and the burned gas
stream discharges from the duct into a region where the pressure is 15 psia,
plot the isentropic stagnation pressure at the duct inlet section against the
Mach Number of the air at the inlet section for fuel-air ratios of 0.02 and 0.04.
To simplify the computations, it is agreed to neglect friction, the change in the
mass rate of flow, the change in chemical composition, and to assume that there
is no diffuser a t the exit of the duct.
7.6. Air is brought into a tube through a converging-diverging nozzle d e
signed for a Mach Number of 2, and has initially a stagnation temperature of
500°R and a stagnation pressure of 100 psia. The air is to be cooled by some
mechanism not involving friction.
(a) Assuming that there are no shocks, compute the maximum possible amount
of heat rejection per pound of air passing through the duct. Find, for this limit-
ing condition, the final temperature, pressure, and stagnation pressure.
(b) Investigate the types of flow patterns which might exist in the nozzle
duct system if the amount of cooling is greater than that computed for part (a).
7.7. A wind tunnel nozzle for a Mach Number of 2 is to be supplied with air
at 30 psia and 80°F. It is desired to estimate the maximum relative humidity
at the nozzle entrance for each of the following conditions:
(a) At no point in the nozzle is a condition of saturated water vapor reached.
(b) Precipitation of water vapor occurs a t no point in the nozzle. It will be
assumed for this purpose that condensation does not occur until the stream
temperature is 110°F below the temperature a t which saturation is first reached.
(c) The amount of water vapor is such that the Mach Number just down-
stream of the moisture condensation shock is exactly unity, assuming that
Ch.7 PROBLEMS 215
7.8. The sketch shows a tube rocket motor, comprising a long tube which is
fed with liquid fuel and liquid oxidant at one end, with hot gases leaving the
open end.
In one test of such a motor, the measured thrust, P, was 676 lb; the pressure
on the back face of the motor, pt,, was
265 psia; and atmospheric pressure, p,,
was 15 psia.
With the fuel-oxidant combination
used, the gases are characterized by
k = 1.3 and R = 30 ft lbf/lbm F.
Because of uncertainties as to PROB. 7.8.
whether combustion is completed and
as to the degree of dissociation, the stagnation temperature of the leaving gases
is not known, but it is suspected to be between 3000°F and 5000°F.
(a) What is the pressure p, in the exit plane, in psia?
(b) What is the average shearing stress acting on the inside walls of the motor,
in psi?
7.9. It has been suggested that the combustion in rocket motors should be
carried out a t high speed in a tube of constant cross-section, as indicated in the
sketch. It is desired to compare the specific thrust (thrust force per unit mass
rate of flow, in lbf sec/slug) of such a rocket motor with the specific thrust of
the conventional motor, also shown in the sketch.
Combusfion
Lomaustion
PROB. 7.9.
corresponding densities are p l and pz. At any radius r the pressure is p and the
velocity is V'. At r = m, the pressure is pat,.
The speed of propagation V , is extremely low (i.e., a few feet per second).
It is, therefore, permissible to assume that the fractional variations in pressure
are so small that the density can be taken as constant except across the flame
front, where there is a large temperature change.
Show that the curve of p versus r is as shown in the sketch and that outside
the flame front it is represented by
GENERALIZED ONE-DIMENSIONAL
CONTINUOUS FLOW
A I 2oniro1 I P +dp
P I Surface I T +dT
T 1 I
F I ~8.1.
. Control surfaces defined for several methods of gas injection and liquid
evaporation.
traveling along with the stream are taken to be the result of external
heat exchange to or from the main stream. As the flows d w and ~ dw,,
thoroughly mixed with the main stream, pass out of the control surface,
they are assumed to be a t the temperature of the latter.
The energy equation for the flow through the control surface of Fig.
8.1 may be written, assuming gravity effects to he negligible, as
w(d& - dWz) = [w(h + dh) + h , ~dw, + hv d w ~ l
The enthalpy change of the main gas stream, dh, is the sum of the
changes due t o chemical reaction and to temperature change. Thus
where
- [h.
hg~ + T] = h . ~- hog = CpdT - To,), (8.12)
and hogand To, are, respectively, the stagnation enthalpy and stagnation
temperature of the injected gas stream.
Inserting Eqs. 8.11 and 8.12 into 8.10, and rearranging, the energy
equation is finally put into the convenient form
the main gas stream, and (iii) the component of body or gravity forces
acting on the material within the control surface in the direction opposite
to that of the velocity vector; V,' is the forward component of the
velocity V gwith which the injected gas dw, crosses the control surface,
and similarly for VL1.
The wall shearing stress is related to the coefficient of friction, f,
though the definition of the latter:
Substituting Eqs. 8.18, 8.19, and 8.20 into Eq. 8.17, and noting that
PV - kP M', we obtain, following rearrangement,
dp k b ~ ' dv2 dw
-+-- + k ~ ~ -( y)-
1 =0
P 2 v2 2 W
(8.21)
where
dw
--
W
dw,
- -+-
W
d w ~
w
ds dT k - l d p
(8.25)
CP T k P
More important, however, is the total entropy change of the main
stream plus injected gas and evaporated liquid. It may be shown that
for unchanged chemical composition
where dS is the total entropy change per unit mass of main gas stream.
These formulas are derived by writing the energy equation and con-
tinuity equation for two sections (1 and 2) of the flow, and by introducing
TABLE 8.1
I NFLUENCE COEFFICIENTS FOR VARIABLE SPECIFIC HEAT AND MOLECUWR WEIGHT
+
+
I + ~ M * ) ( I+ k7
-l
M (I M ) 2, M*)
1 - M= 1 - M'
WCP
(Note 2)
N OTES : (1) Each influence coefficient represents the partial derivative of the variable in the left-hand column with respect to the variable in the top
row: for example
(2) For unaltered chemical oom~ositiononly, and referring to entropy change of main stream. See Eq.8.26 for total entropy change.
Art. 8.4 FLOW WITH CONSTANT cp A N D W 229
(8.33)
or, in differential form, as
k-1
-M2
dTo - dT
To T
+ 2
k-1
dhP
M2
(8.34)
,, ,
or, in differential form,
dM2
2 (1 + k--1
2 M a ) dA (1 + kM2) (1 + IC-I
2 M') dTo
I
-
NOTE: Each influence coefficient represents the partial derivative of the variable in the left-hand column with respect to thevariable in the top row; for
exam~le
232 GENERALIZED ONEDIMENSIONAL FLOW Ch. 8
USEFUL INTEGRAL RELATIONS. The several useful integral relations
between the properties a t two cross-sections may be summarized as
yM,j
k-1 k--l
I+- h2
Po2
- P2
=-
2
P OI Pl +
from which it may be shown that both jdx and d X are always positive,
i.e., negative friction or negative drag would constitute violations of the
Second Law.
234 GENERALIZED ONEDIMENSIONAL FLOW Ch. 8
Considering next the case where there is no chemical change, no fric-
tion, no drag, and no W,, we find from Eq. 8.26 that
and that the values of Az, (Q - W , + AH)1-2,2 2 - XI, WZ, W2, k2,
etc. are known for a section 2, a small distance downstream of section 1.
Then, from Eq. 8.27, the value of M2 may be found by numerical or
graphical integration. Next, Eq. 8.28 is solved for T2, Eq. 8.29 for p2,
and Eqs. 8.30 and 8.31 for V2 and p2, respectively. The procedure is
then repeated for a step between sections 2 and 3, and so on.
Integration Near Mach Number Unity. Special precautions in numer-
ical integration must be taken near M = 1because many of the influence
coefficients there approach infinity. It is usually possible to avoid diffi-
culty by obtaining an approximate analytical formula for a step near
M = 1, making use of the approximation that 1 - M2 is negligible com-
pared with unity. For example, consider two of the terms of Eq. 8.40:
Integration yields
I C+l
A* MVL k+l J
0
I;:
240 GENERALIZED ONE-DIMENSION.4L FLOW Ch. 8
so that
When the flow is subsonic, the adiabatic wall temperature Tawdoes not
differ very much from the stagnation temperature To. Moreover, the
error in substituting To for Tawwill be small in any event if Twis con-
siderably in excess of Taw. We shall assume that this substitution
(which simplifies the subsequent calculations) is permissible; this is
equivalent to assuming that the recovery factor is unity. Then the
foregoing equation yields
dTn 4% dx
has an accuracy of a few per cent for fully developed turbulent gas flows.
Substituting this formula into the preceding equation, we get
With the help of Eq. 8.52, we may now write Eq. 8.48 as
The procedure beyond this point depends on the nature of the heat-
transfer process. To illustrate, we shall consider three specific problems:
(i) constant wall temperature, (ii) constant heat flux, and (iii) nuclear
reactors.
Case of Constant Wall Temperature. The wall temperature is ap-
proximately constant when the tube metal is a good c o : ~ J ~ ~ cand
t t , r is
244 GENERALIZED ONE-DIMENSIONAL FLOW Ch. 8
Hence, for any initial value of Tol, and for known values off, D, and
T,, Eq. 8.54 yields Toz as a function of x2.
NUMERICAL INTEGRATION FOR MACH NUMBER VARIATION. TO com-
pute the variation of M with distance, we must integrate Eq. 8.53. In
the present example it is, unfortunately, not, possible to separate variables
and integrate in closed analytic form. Hence it is necessary to integrate
either numerically or graphically. We shall first illustrate numerical
integration.
Suppose that f = 0.005, M I = 0.5, and TWITol= 4. Let us specify
section 2 as that section downstream of section 1 where To2/TO1 = 1.05.
The location of section 2 is found by direct substitution into Eq. 8.54,
thus yielding
22 - 2 1 1 4- 1
-- - - -In = 1.68
D 0.01 4 - 1.05
In other words, section 2 is 1.68 duct diameters downstream of section 1.
Eq. 8.53 is now integrated approximately in finite-difference form
over the short interval between sections 1 and 2, the approximation
being that the coefficient of dTo is constant (at its mean value) during
the interval. Let P Tand~ Ff denote the values of the influence coeffi-
cients when evaluated a t &i = ( M I + M2)/2. Furthermore, let us set
+
To = (Tol T o 2 ) / 2for the interval. Then the approximate integration
Trial 111: After plotting for Trials I and I1 the computed values of M2
against the guessed values, we try Ma = 0.5293 and find that this value
makes the solution converge with adequate accuracy.
The corresponding values of all other properties a t section 2 may be
found either through numerical integration of the corresponding differ-
ential equations (as given in Table 8.2) or, more conveniently, through the
integral relations of Eqs. 8.41 to 8.46. The latter method is used here.
From Eq. 8.45,
b - 1
FIG.8.4. Illustrative results of calculations for combined friction and heat transfer
in a constant-area tube, with k = 1.4, (R = 1, and MI = 0.5. Solid curves are for
constant wall temperature and dashed curves are for constant heat flux per unit area.
The circled points of each curve indicate choking (M = 1).
(a) Stagnation temperature versus distance.
(b) Mach Number versus distance.
(c) Stagnation pressure versus distance.
(d) Temperature versus distance. Note that just prior to choking the tempera-
ture decreases as heat is added.
Art. 8.9 COMBINED FRICTION A N D HEAT TRANSFER 247
which again allows us to use the tabulated numerical values in Table B.2.
Returning now to our illustrative problem, the changes which occur
between section 2 and a section 3 further downstream are determined
in the same manner as the changes between sections 1 and 2. This
procedure is continued until the exit of the duct is reached, unless Mach
Number unity is reached first. The latter result indicates choking, and
248 GENERALIZED ONE-DIMENSIONAL FLOW Ch. 8
then either the entrance Mach Number or the length of duct must be
reduced in order to obtain a steady-state solution.
The accuracy of the numerical integration naturally depends on the
size of interval chosen for each step and is improved by the selection of
small steps.
Fig. 8.4 shows the results of numerical calculations for the example
considered here. Also shown in this figure are corresponding results for
the case of constant heat flux per unit length of tube. It is seen that
many of the fluid properties change very rapidly as choking is a p
proached. Fig. 8.4g shows that high rates of heat transfer make the
choking problem very serious. For example, with TW1/Tol% 5 and
f 2 0.005, only about 10 diameters of duct length are allowable when
MI = 0.5. Apart from the large stagnation-pressure losses associated
with high Mach Numbers, therefore, considerations of choking often
prohibit the use of high values of M in heat exchangers.
GRAPHICAL INTEGRATION BY THE METHOD OF ISOCLINES. TO illustrate
this rapid method for obtaining solutions with the accuracy typical of
graphical constructions, we combine Eq. 8.53 with Eq. 8.52 and thus
obtain a differential equation of the form (assuming f is constant)
Now consider a chart with M~as ordinate and 4jx/D as abscissa (Fig.
8.5). Eq. 8.56 gives the slope on this chart of an integral curve satisfying
the differential equation. Consider-
ing the same illustrative problem
as before, we are given a constant
value of T,, and Tol is known.
The value of To corresponding to
any value of 4jx/D may, therefore,
be computed immediately from
Eq. 8.54. Hence T, and To are
known a t each value of 4fx/D.
If, for a given value of 4jx/D,
we choose various values of M~at
convenient intervals, the corre-
Oo
L sponding slopes may be found
from Eq. 8.56. These slopes are
4fx/D plotted in the chart as short
Fxa. 8.5. Illustrates graphical integra- streaks. By repeating this pro-
tion by method of isoclines.
cedure for a greatmany values of
4fx/D, the entire chart may be filled with such streaks indicative of
the local slopes.
Art. 8.9 COMBINED FRICTION AND HEAT TRANSFER 249
advantage of being in closed form and which, therefore, does not require
the tedious integrations described above.
Our interest lies primarily in the stagnation-pressure loss and the
increase in stagnation temperature occasioned by given values of T,,
pol, Tol, T,, D, and L. The changes in stagnation temperature are
easily calculable from Eq. 8.54 and need no further consideration.
For the variation in stagnation pressure we find from Table 8.2, for
the case under consideration, that
Eliminating dTo with the aid of Eq. 8.52, we get, in dimensionless form,
By eliminating p2/pl from Eqs. 8.42 and 8.46, and setting w2/w1 = 1
and A2/A1 = 1, we obtain k+l
k-1 k-1
(8.58)
Substitution of this relation for M2 into Eq. 8.57 then yields the differ-
ential equation
250 GENERALIZED ONE-DIMENSIONAL FLOW Ch. 8
where
and
The quantity (13 depends on MI and M2, and may be set equal to unity
as a first approximation. The definite integral D depends on the varia-
tion of T , and of To with x; this integral, accordingly, depends in part on
the method of heating.
As shown previously, and in accord with Eq. 8.58,
Assuming that T,, Tol, L, D, and pol are known t o begin with, and
that we wish to calculate p02 and M2 corresponding to a certain inlet
Mach Number MI, we proceed as follows: The value of a)is found from
Eq. 8.62, which is derived below. A first approximation for po2is found
from Eq. 8.59a by setting @ = 1 (or perhaps a slightly higher value,
depending on one's experience with prior calculations). The corre-
sponding value of M2 may then be found from Eq. 8.60, inasmuch as T02
is easily calculated from Eq. 8.54.
If it is desired to improve on this approximation (often i t is not neces-
sary), we employ the first approximations to po2 and M2 for calculating
a more accurate estimate of 03 from Eq. 8.61. The calculation procedure
is then repeated with this new value of 03. The recalculation of and
repetition of the calculation procedure are continued until convergence
is obtained; in practice convergence is usually adequate on the second
approximation.
In order to evaluate a), we first write Eq. 8.54 aa
Case of Constant Heat Flux. We now take up briefly the case where
the heat flux per unit area is the same for all values of x. Such a situation
results, for example, when the tube is heated electrically by passing
current either through the wall of the tube itself or through resistance
wires wrapped uniformly around the tube.
252 GENERALIZED ONE-DIMENSIONAL FLOW Ch. 8
value at the ends. Assuming that the tube has a length L, and employ-
ing the dimensionless length variable t = x / L , the equation representing
such a parabolic distribution of heat input F
dt max
The three ratios on the right-hand side are now expressed in terms of E
by means of Eqs. 8.67c, 8.67d, and 8.68, thus yielding
equal to zero, we find the following value of &, for which the wall tempera-
ture is a maximum:
M < 1 M = l M > 1
G<O
G=0
-
0
m
indeterminate
+0
G>O + m -
Therefore the curve of M versus x may have many different forms, de-
pending on whether the initial Mach Number (MI) is less than or greater
than unity, and also depending on whether G is always positive, always
negative, or changes sign. We shall now explore the various possibilities.
CASE I: G IS ALWAYS NEGATIVE. (Fig. 8.7a.) If the flow is initially
subsonic, the Mach Number continually decreases. If it is initially
supersonic, on the other hand, the Mach Number continually increases.
(a) G < 0
I
X
- x X
(d) G Changes from Negative to Positive (e) G Changes from Positive to Negative
FIG. 8.7. Illustrates possible variation s of Mach Number versus distance.
In brief, the Mach Number always proceeds away from unity. There
are no limits on the flow.
Examples: (i) Isentropic flow in a diverging passage; (ii) Flow with
cooling only; (iii) Adiabatic flow with friction in a diverging passage,
with d(1n A)/dx > 2 k ~ ~ f / D .
Art. 8.10 SPECIAL CONDITIONS A T THE SONIC POINT 257
Inserting Eq. 8.75a into Eq. 8.74, and solving the resulting quadratic
for (dM2/dx)*, we finally obtain
PROBLEMS
8.1. Plot on a temperature-entropy diagram the curves corresponding to
Simple Friction, Simple To-Change, and Simple Gas Injection with y = 0.
Choose the starting point of each curve such that in all the processes the cor-
responding stream properties are identical at M = 2.
8.2. A stream flowing in an insulated duct with friction is to be maintained
at constant Mach Number through suitable changes in duct area.
Assume that a one-dimensional treatment is acceptable and that at section
1 of the duct the properties are MI,pl, Al, etc.
Art. 8.10 PROBLEMS 261
(a) Starting with fundamental principles, show that the product of area and
pressure is the same for all cross sections.
(b) Find an expression for the area a t a point downstream of section 1 in
terms of the area a t 1, the Mach Number, the friction coefficient, anti the
number of length-diameter ratios (based on Dl) between section 1 and the
downstream section.
(c) If MI = 0.5, pl = 1 atm, and f = 0.005, compute the ratios A2/Al, p2/pl,
and poz/pol if section 2 is 50 diameters downstream of section 1.
8.3. Air flows adiabatically in a tube of circular cross section with an initial
Mach Number of 0.5, temperature of 100O0R, and pressure of 100 psia, all a t
section 1.
The tube is to be changed in cross-sectional area so that, taking friction into
account, there is no change in the temperature of the stream.
Assuming the friction coefficient to be 0.005 and that the exit section (2) is
downstream of the inlet by a distance 100 times as large as the initial diameter,
find :
(a) The final Mach Number, Mz
(b) The ratio of diameters, Dz/Dl
(c) The find stagnation pressure, p02, in psia
8.4. A hydrocarbon gas is to be injected step-wise and burned in a high-
speed stream of air. It is desired to carry out the combustion a t a constant
Mach Number of 0.5, and changes in cross-sectional area will therefore be
necessary.
To simplify the analysis, a one-dimensional point of view will be adopted,
and the following assumptions will be made: (a) variations in specific heat and
molecular weight are negligible, (b) wall friction is negligible, (c) the hydrocarbon
is injected a t low speed, and (d) enthalpy effects associated with the difference
between the gas temperature and the stream temperature are negligible com-
pared with the enthalpy effects due to combustion.
Assuming that the constant-pressure heat of reaction is 15,000 Btu per pound
of hydrocarbon gas, Tol = 1000°R, and Toz = 250OoR, estimate the area ratio
A2/A1 and the pressure ratio pz/p~.
8.5. A gas flows in a conical duct of circular cross section, the included angle
between the walls being 8. If j is the coefficient of friction, find a relation be-
tween 0, DID*,M, and f, where D is the diameter a t Mach Number M and D* is
the diameter a t Mach Number unity. Find a similar relation between pO/po*,
M, 8, and f.
NOMENCLATURE
acceleration absolute temperature
area velocity component in xdirec-
area vector tion
speed of sound velocity component in ydirec-
force tion
acceleration of gravity velocity
enthalpy per unit mass velocity vector
unit vectors in the x-, y-, and velocity component in zdirec-
zdirections respectively tion; mass rate of flow
ratio of specific heats Cartesian coordinates
length
mass angle
curvilinear coordinate meas- circulation
ured normal to streamline angle
pressure density
rotation vector velocity potential
radius; cylindrical coordinate stream function
radius vector rotation; cylindrical coordi-
radius of curvature of stream- nate
line angular velocity
entropy per unit mass
time signifies stagnation state
where V denotes the vector velocity - and r the vector radius from any-
fixed origin. If u, v, and w denote respectively the velocity components
in the x-, y-, and z-directions, the Cartesian
form of Eqs. 9.la and 9.lb becomes
dx
I
I
dx
I - '"
- CIRCULATION PER UNIT AREA. Let us now
consider the circulation d r , around a small
FIG. 9.2. Circulation square element in the x,y-plane, as shown in
f o r elementary
curve.
Fig. 9.2. Employing Eq. 9.lc with the velocity
components shown in Fig. 9.2, and proceeding
counter-clockwise from the origin, we obtain
dr,=udx+
or, simplifying,
Similar expressions for the other directions may be found b y the usual
rotation of indices.
GREEN'STHEOREM. Considering now the circulation around an area
of finite size, we note from Fig. 9.3 that the line integral around the
bounding curve C is the algebraic sum of the line integrals around the
elementary square elements comprising the area
within C, since each interior line of each ele-
ment is traversed an equal number of times in
opposite directions, and hence only the exterior
lines of the elements (which comprise curve C)
contribute to the circulation. Therefore, em-
ploying Eq. 9.2, we find that the circulation
around any finite closed curve C in the x,y- FIG. 9.3. The circuia-
plane may in the limit be expressed in terms tion around any closed
of a surface integral, namely, curve is the sum of the cir-
culations around all the
closed curves bounding
r, =
f(u~x+u~Y)
A comparison between Eqs. 9.4 and 9.5 shows that taken together they
represent Green's Theorem for the transformation from line to surface
integrals in two-dimensional space.
STOKES' THEOREM. The general form of Eq. 9.4, expressed in vector
notation, is,
'1
tion OA', and the relative vertical
gdydt displacement AA' is
av
AA' = - dx dl
ax
Hence the angle AOA', taken posi-
tive when measured counter-clock-
=dl gdxdl
U
X
wise, is given by
dx
v
au
LAOA' = -dt
F IG . 9.4. Illustrates fluid rotation at
ax
a point. OA and OB are mutually per-
pendicular fluid lines. and the time rate of change of this
-
angle is avlax.
Similarly, it may be shown that the angular velocity of the fluid line
OB is - au/ay.
From the definition of rotation, therefore,
Referring to Eq. 9.3, we see that the circulation per unit area is twice
the average rotation of a fluid particle:
and, referring to Eq. 9.8, we conclude that the angular velocity fi of the
entire mass is, as we expect, identical with the average rotation of each
particle.
EXAMPLE OF IRROTATIONAL MOTION. A simple but important exam-
ple of irrotational motion is the potential vortex, which describes approxi-
mately the motion in tornadoes and whirlpools, and which plays an
important role in the theory of flight and the theory of turbines and
compressors. In its simplest form a potential vortex is a two-dimensional
272 EQUATIONS OF STEADY, IRROTATIONAL FLOW Ch. 9
motion in which the streamlines form concentric circles, and the tan-
gential velocity along any streamline is inversely proportional to the
radius of the streamline. Thus,
Thus the circulation of any element not enclosing the origin is zero, and,
employing the theorem illustrated by Fig. 9.3, we conclude that the
circulation of any area not enclos-
ing the origin of the vortex is zero.
Considering next the circulation
around the streamline of radius r
in Fig. 9.6, we obtain
2r
r =lZ1VrdB = K l d0 = 2rK
The line on the cork will in fact have the same type of rectilinear motion
as the magnetic needle of a compass when the compass travels through
a circular path in a horizontal plane.
Dw
F , = ma, = m-
Dt
274 EQUATIONS OF STEADY, IRROTATIONAL FLOW Ch. 9
where F,, F,, and F, denote the respective forces acting on the system
in the x-, y-, and z-directions, and u, v, and w are the corresponding
components of velocity. The symbol DjDt represents the substantial
derivative with respect to time and denotes that the differentiation is to
be carried out while following a particular fluid particle.
Since shearing forces and body forces are excluded from this analysis,
the only forces acting on the system are pressure-area forces. The
magnitudes of these forces acting on the various faces in the x,z- and
+
y,z-planes are shown in Figs. 9.7a and 9.7b a t the times t and t dt,
respectively.
Considering first the x-direction, the average force during the time
interval dt is
The mass of the system is the product of the mass density and the
volume:
m = p 6% 6y 62
The acceleration in the x-direction, a,, may be written
Art. 9.3 EULER'S EQUATIONS OF MOTION 275
Now, since dx is the x-displacement of the fluid particle during the time
dt, it follows that u = dxjdt. Likewise, v = dy/dt and w = dz/dt
Hence
Inserting the foregoing expressions for m, for F,, F,, and F,, and for
a,, a,, and a , into Eqs. 9.12, and simplifying, we finally obtain Euler's
equations of motion in Cartesian coordinates:
There are two special sets of conditions for which this system of equa-
tions may be easily integrated: first, along a streamline; and second,
throughout the entire flow field, provided that the flow is irrotational.
Integration of Euler Equation Along a Streamline in Steady Flow.
Since a streamline is defined by
Y
the condition that a t each instant
the velocity vector is tangent t o
the streamline (Fig. 9.8), we may
write the equation
- of the stream-
line in the form
dy -d =z- . w d- x =u-
_ = -v.
dx u' dy v ' dz w
au au au
du = - dx
ax
-t - d y
ay
+-
az
dz; etc.
we obtain
where V is the resultant velocity. Eq. 9.15 may be put in the form
J v 2
+ = constant along streamline (9.16b)
the assumptions given here, valid only from point to point of a given
streamline. Or, to put it differently, the Bernoulli constant may in
general vary from one streamline to another. Of course, if all the stream-
lines originated in a region of uniform flow, the same Bernoulli constant
would prevail throughout the entire field of flow.
Integration of Euler Equation for Steady, Irrotational Flow. We now
abandon the restriction that only points along the same streamline are
to be considered and introduce instead the assumption that the motion
is irrotational and steady. According to Eq. 9.7, then,
the last two equations being obtained from the first through rotation of
indices.
We now multiply Eq. 9.13a by dx, Eq. 9.13b by dy, and 9 . 1 3 ~by dz,
and introduce the assumption of irrotationality through Eq. 9.17, thus
getting
that
a + v2 + w 2
4 )= t ( ~
etc. ) *
u2
ax 2
278 EQUATIONS OF STEADY, IRROTATIONAL FLOW . Ch.9
and that
Dt
Now, from Eq. 9.13a, we note that in the absence of friction,
Carrying out similar operations on the other two terms of the line
integral, we obtain
Now, if the density is a function only of the pressure, it is clear that the
line integral depends only on the fluid properties at points A and B, and
not on the path of integration. Points A and B coincide if the fluid
line is closed, and then the line integral is zero. The proof of the theorem
is completed by noting that the line integral of the velocity around a
closed curve is by definition the circulation.
280 EQUATIONS OF STEADY, IRROTATIONAL FLOW Ch.Y
When body forces, such as gravity, are present, Kelvin's Theorem re-
mains valid provided that the body force may be written as the gradient
of a potential. such body forces are usually said to be conservative.
IRROTATI~NAL MOTION. If the fluid starts from rest, or if in some
region the flow is uniform and parallel, the rotation in these regions is
zero. Kelvin's theorem then leads to the important conclusion that,
the entire flow is irrotational, subject to the assumptions of no friction,
a homogeneous fluid, the existence of potentials for the body forces, and
the absence of discontinuities.
Since in many practical problems the flow may as a first approxima-
tion be thought of as frictionless and as having started as a uniform,
parallel flow, Kelvin's theorem and the concept of irrotational motion
play an important role in the understanding of many flow phenomena.
They indicate, for example, that while in most flow patterns observed
in nature the fluid particles follow curved paths, the flow is nevertheless
likely to be irrotational except in those regions where viscous stresses are
an important factor.
dr = V R d0 - (R + dn) dB = -
Finally, after equating the force to the product of mass and accelera-
tion, and simplifying, we get the steady-state form of Euler's equation
in a direction normal to the streamline direction:
1
T ds = dh - - dp
P
may be expressed in derivative form as
ah as 1 ap
--
- T-+--
an an pan
Substituting this into the expression for the rotation, we obtain finally
a special form of Crocco's Theorem for two-dimensional flows:
{ypzld t
areas of the control surface, i t is pvdxdzt
seen that the net flow into the con- d p~d~dz)dy
a
-- pvdxdz
(pu dy dz) dx
ax *
=
a
-- (pu) dx dy dz
x
FIG.9.11. M a s flows through elemen-
ax tary control surface.
Likewise, the net flows into the control surface in the y- and z-directions
only are, respectively,
and
- -a (pw) dz dx dy
az
The time rate of change of mass within the control surface is expressed by
a ap
- (p dx dy dz) = - dx dy dz
at at
Hence the principle of conservation of mass takes the general form
Although, in the previous derivation, use was made of the First Law of
Thermodynamics in the form of the steady-flow energy equation, Eq.
9.27 may be obtained for isentropic flows with no recourse to thermo-
dynamics except the p-p relation of Eq. 9.26. The reader may verify
that Eq. 9.27 may be derived from the Euler equation (Eq. 3.18), the
definition of the sound velocity (Eq. 3.3), and the isentropic pressure-
density relation for a perfect gas (Eq. 9.26). The steady-flow energy
equation in reality embodies Newton's second law in the form of the
kinetic-energy term, and it is only the latter term which entered into
the previous derivation of Eq. 9.27.
Thermodynamics plays a more important role, it will be seen later,
when shock waves, friction, heat transfer, or work effects are important
factors in the flow pattern.
it follows that the line integral of the velocity between any two points
A and B depends only on the location of the points and not on the path
of integration. For, referring to Fig. 9.12 (see also Fig. 9.1), we may
write for the closed curves AlB3A and A2B3A :
0
lBv cos a! dl = lBv
0
00s a! dl
points in a n irrotational
motion is independent of
path of integration.
Since paths 1 and 2 were chosen arbitrarily, it follows that the line
integral of the velocity is independent of the path.
This means that V cos a! dl is an exact differential, and may, therefore,
be considered as the differential of a point function whose value depends
only on x, y, and z. This function is called the velocity potential, p, and
is defined through the relation
dp = V cos a! dl
d~ =
- V cos a!
dl
Thus, the derivative of the velocity potential in a given direction repre-
sents the component of velocity in the same direction. In Cartesian
coordinates we have
u = ap/az = p, (9.29a)
V = grad p = Vp
where the last term is set equal to zero because the cross product of two
parallel vectors is zero.
The concept of the velocity potential derives historically from the
concept in mechanics of the potential of a force. It may be recalled
that for conservative systems, the work, or the line integral of the force,
is zero around a closed curve, and hence a force potential exists whose
derivative in any direction is the force in that direction.
Because a velocity potential always exists for irrotational motion,
the terms potential motion and irrotational motion are often used inter-
changeably.
or, expanding,
Art. 9.8 DIFFERENTIAL EQUATIONS-VELOCITY POTENTIAL 287
,.
actual pressure and density changes. Thus, c2 = dp/dp. Introducing
Eq. 9.31, we obtain
dp = -dp = - - dP
2
(a2 + + .:)
c cZ
The local sound velocity is, of course, not constant, but varies with p
according to Eq. 9.27:
Substitution of Eq. 9.33 into Eq. 9.32 would yield a single differential
equation for p in terms of x, y, and z, an equation which simultaneously
satisfies the law of conservation of mass, Newton's second principle of
motion, and the laws of thermodynamics.
288 EQUATIONS OF STEADY, IRROTATIONAL FLOW Ch. 9
DIRECT AND INVERSE METHODS. Having obtained Eqs. 9.32 and 9.33,
how do we employ them for the solution of problems? Two approaches
are possible:
(i) Most often, we are interested in the flow through a duct or around
a body of known shape, and with certain additional boundary conditions
as to velocity and pressure. We must, therefore, seek a function p(x, y, Z)
which satisfies Eq. 9.32 and which also satisfies all the boundary condi-
tions. This is often difficult, if not impossible, and experience and
intuition rather than a prescribed procedure are likely to lead to success-
ful results. Thus, although for practical reasons the direct approach
is preferable, it is often necessary to use the second approach.
(ii) I n this case we find, more or less a t random, functions P(X, y, Z )
which satisfy the differential equation, Eq. 9.32, and then proceed to
determine the corresponding streamline patterns. If one or more of the
streamlines coincides approximately with the shape of a solid body or
duct which is of interest, then the solution is useful; otherwise, it is not.
Evidently, this procedure is at best haphazard.
In either case, if we have a solution of Eq. 9.32 for p in terms of x, y,
and z , the entire flow pattern may then be determined. From Eq. 9.29
it is possible to calculate the three velocity components and the Mach
Number a t each point and thus to construct the streamline pattern.
Furthermore, from the equations for isentropic flow, Eqs. 4.14a, 4.14b,
and 4.14c, all the fluid properties may be found a t each point. The
integration of the pressure-area forces on the solid boundaries then
gives the resultant force acting on the duct or on the body immersed
in the stream.
MATHEMATICAL REMARKS. Mathematically speaking, Eq. 9.32 is of
the second order. A differential equation is said to be linear when the
dependent variable and its derivatives appear only in linear form. In
Eq. 9.32, however, we find quadratic terms of the form e2, and cubic
terms of the forms pZ2p,, and rp,p,cp,,, which means that Eq. 9.32 is
nonlinear. An important feature of linear equations is that solutions
may be superposed; that is, if +y(x, y, z) and p2(x, y, Z) are respective
solutions of a linear differential equation, then Alpl + A2pz is also a
solution, where A1 and Az are arbitrary constants. Many of the mathe-
matical difficulties encountered in compressible fluid mechanics are
traceable to the fact that the equations are nonlinear, for i t is not
possible to build up complex solutions by superposition of a few simple,
known solutions. Later, it will be seen that, a t the expense of other com-
plications, the hodograph method removes this difficulty, for the differen-
tial equations prove to be linear when the velocity components rather
than the physical coordinates are used as the independent variables.
Afi. 9.8 DIFFERENTIAL EQUATIONS-VELOCITY POTENTIAL 289
Definition oj (o.
Equation oj continuity.
a a a
- bur)
ar
+-
am (PV)+ r - (pw)
a2
= 0
The combination of Eqs. 9.40 to 9.43 in the same manner as that used
previously for Cartesian coordinates yields finally
where I
Art. 9.9 DIFFERENTIAL EQUATIONS-STREAM FUNCTION 291
+
If is a point function, the order of differentiation is immaterial. By
comparison of the foregoing equation with Eq. 9.50 we see that the
+
statement that there exists a point function defined through Eq. 9.51
is equivalent to the statement of the continuity equation. Conversely,
the continuity equation is the necessary and sufficient condition that +
be a point function.
The magnitude of the velocity vector a t any point is given in terms of
the stream function by the relation
Now, referring to Fig. 9.14, the mass rate of flow dw across any surface
AB (of unit width normal to the paper) connecting two neighboring
surfaces of constant +
is given by
dw = pu dy - pv dx
Y and hence we see that the change
+
in the magnitude of between two
+
lines of constant is a measure of
the mass rate of flow between the
+
two lines, according to the rela-
tions
d w = po d+ (9.54a)
FIG.9.14. Illustrates physical signif-
icance of stream function. w = PO(+B - # A ) (9.54b)
Since, for steady flow, the mass rate of flow between streamlines is
+
constant, it follows that lines of constant are streamlines, hence the
term "stream function."
The Differential Equation of the Stream Function. We shall now
derive a single differential equation representing the steady, two-dimen-
sional, irrotational, frictionless, isentropic flow of a perfect gas.
Art. 9.9 DIFFERENTIAL EQUATIONSSTREAM FUNCTION 293
From the Euler equation and the expression for the sound velocity
in isentropic flow, we obtain
Eqs. 9.57, 9.58, and 9.59 on the one hand are completely equivalent
to Eqs. 9.34 and 9.35 on the other. However, the differential equations
in terms of q are simpler in form than those in terms of J., and hence it is
customary to work with Eq. 9.34.
INCOMPRESSIBLE FLOW. For incompressible flow, the sound velocity
is infinite relative to the velocity of the gas, and Eq. 9.57 reduces to
294 EQUATIONS OF STEADY, IRROTATIONAL FLOW Ch. 9
This is the Laplace equation in two dimensions, and because of its
similarity in form with Eq. 9.37, a solution for the potential lines in plane,
incompressible flow represents also a solution for the streamlines, and
vice versa.
Axially Symmetric Flow. Since in steady, axially symmetric flow
there are only two space coordinates, the statement of the continuity
equation is the necessary and sufficient condition for the existence of a
stream function. Bearing in mind that partial derivatives with respect
to w are zero, Eq. 9.41 takes the form
a a
-
ar
(pur) +r-
az
(pw) = 0
Using the Euler equation and the expression for the sound velocity
in isentropic flow, we get
Setting d$ equal to zero and using Eqs. 9.68 and 9.69, we find the slopo
of a line of constant J, to be given by
Using Eqs. 9.68 and 9.69, we get for the slope of an equipotential line,
Since p/po cannot be greater than unity, i t follows that the potential
lines are more closely spaced than the streamlines. And, since p/po
is near unity a t low Mach Numbers and far less than unity a t high Mach
Numbers, the difference between the spacings will be most marked in
Art. 9.10 VELOCITY POTENTIAL AND STREAM FUNCTION 297
PROBLEMS
9.1. Show that Euler's equation for unsteady, irrotational motion can, with
the introduction of the velocity potential q,be integrated to give
9.2. I n Art. 9.3 Euler's equations in Cartesian coordinates (Eq. 9.13) were
obtained by applying Newton's second law of motion to a fluid particle of
fixed mass and identity, and with sides 82, 6 y , 82.
Show that the same results may be derived by applying the momentum theo-
rem to the flow through a cubical control surface fixed in space and with sides
ax,d y , dz.
9.3. I n Art. 9.6 the equation of continuity (Eq. 9.23) in Cartesian coordinates
was derived by applying the principle of conservation of mass to conditions
within and a t the boundaries of a control surface fixed in space and with sides
ax, d y , dz.
Show that the same result may be obtained by applying the principle of con-
servation of mass to the motion of a particle of fixed mass and identity and with
sides 62, 6y, 62.
9.4. Starting with first principles, verify Eqs. 9.39 through 9.45, expressing
the steady, irrotational, frictionless, isentropic motion of a perfect gas in terms
of the velocity potential and in terms of cylindrical coordinates.
9.7. Show that the stream function for axially symmetric flow, as defined
by Eq. 9.62, defines surfaces which are generated by streamlines. Derive the
+
relation between the magnitudes of between two streamline surfaces and the
corresponding mass rate of flow between these surfaces.
Ch.9 PROBLEMS 299
9.8. Consider axially symmetric flow, with the streamlines lying in planes
containing the axis of symmetry. Use the following notation for a spherical
aystem of coordinates:
R radius vector, measured from the origin 0
cr angle of the radius vector with respect to the axis of symmetry
u velocity component along radius vector
u velocity component perpendicular to radius vector
represents in polar coordinates a plane source motion for which w is the maas
flow issuing from the line source per unit length of source.
Assuming the flow is frictionless and adiabatic, calculate the corresponding
expression for the velocity potential for
(a) Incompressible flow
(b) Compressible flow with a stagnation sound velocity co
9.10. Show that the potential function
NOMENCLATURE
angle of attack V velocity
area x, y Cartesian coordinates
sonic speed
specific heat a t constant pressure dl - h L 2
coefficient of lift camber ratio
moment coefficient thickness ratio
pressure coefficient interference pressure coefficient
amplitude of wave-shaped wall affine transformation factor
ratio of specific heats mass density
chord length of air foil; also perturbation velocity potential
wave length of wave-shaped velocity potential
wall
a characteristic dimension; also ( )o signifies stagnation state
lift force ( ) signifies free-stream condition
Mach Number ( ) signifies state at solid boundaries
pressure ( ) signifies incompressible flow
gas constant (M, = 0)
absolute temperature
perturbation velocity compo- [0] signifies "of the order of mag-
nent in xdirection nitude of"
perturbation velocity compo-
nent in 21-direction
where cmis the sound velocity a t points where the stream velocity is Vm.
306 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 10
+ 101(Mm2 $)
Vw
+...
together with terms of higher order. We now assume that the perturba-
tion velocities are so small compared with Vw that we may write
"2($)<< 1 2
(10.12)
2
k2(&) << 1 (10.13)
I---
*m2
c2
2 1 - Mm2 - [2Mm2 + (k - 1)Mm4]-
v,
U
(10.15)
Substituting Eqs. 10.5, 10.7, 10.8, 10.15, 10.16, and 10.17 into Eq. 10.1,
we get the approximate differential equation
where
8==- (10.21b)
308 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 10
Examination of Assumptions Underlying Linearized Equations. The
assumptions underlying the linearized differential equation are embodied
in Eqs. 10.12, 10.13, 10.14, 10.19, and 10.20. Examination of these
equations indicates that the latter two include the former three, and are,
therefore, the most stringent.
As M, approaches unity, Eq. 10.19 becomes invalid. Hence, except
for extraordinarily small perturbations, the conclusions set out in the
remainder of the chapter are not valid for transonic flow, i.e., for values
of M, greater than about 0.8.
The perturbation ratios u/Vw and v / V , depend chiefly on the thick-
ness ratio of the body. The accuracy of Eqs. 10.19 and 10.20 will, for
thin profiles, be good up to high subsonic (but not transonic) Mach
Numbers. For thick profiles, on the other hand, the assumptions are
reasonably valid only for very low Mach Numbers.
Since the left-hand side has zero dimensions, the right-hand side must
also be dimensionless. But p,, c,, and L can in no way be formed into
a dimensionless group. From this we conclude that, when the physical
equation is rearranged as indicated, these three terms disappear from
the equation except insofar as pm and c, appear in the dimensionless
groups.
We now solve this expression for T/T,, and, aftcr cmploying the relations
cm2= kRT,
Mm = VJc,
we obtain
310 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 10
We now note that the second-term in the bracket is less than unity, and
so we develop a series expansion for p/pm with the help of the binomial
theorem:
We now introduce the assumptions of Eqs. 10.12, 10.14, and 10.19, and,
since mathematical simplicity is the main purpose of the iinearized
theory, we make the additional assumption that
The final expression for the pressure coefficient of the linearized theory
is then
u
Cp 2 -2- (10.26)
V,
Eq. 10.25 is a more severe restriction than Eq. 10.19. However, in
many applications, Eq. 10.26 for C, is used only when comparing cor-
responding values of C;, for two flows wherein the two values of C, do
not differ by much. Such is the case, for example, in the formulation of
the Gothert and Prandtl-Glauert similarity rules to be discussed later
in this chapter. In cases of this sort, i t may be shown by detailed con-
sideration that the use of Eq. 10.26 in comput,ing the ratio of two values
of Cp involves only the assumption that ( u / V ~ < )<~ I, rather than the
assumption of Eq. 10.25.
5
We shall now investigate a particular solution (I) o the linearized
equations, the practical significance of which rests on the fact t,hat in a
relatively simple way we may demonstrate the quantitative and yualita-
tive effects of compressibility on the streamline pattern and on the pres-
sure distribution.
Art. 10.4 FLOW PAST A WAVE-SHAPED WALL 311
where h and I are length constants whose physical meanings will be evi-
dent later and 13 is defined by Eq. 10.21b. Whether or not this assumed
function satisfies the conditions of continuity, momentum, energy, and
irrotationality depends on whether it satisfies Eq. 10.21a. To determine
this, we first take derivatives as follows:
Substitution of Eqs. 10.29 and 10.31 into Eq. 10.21a shows that the
function cp(x,y)of Eq. 10.27 satisfies the differential Eq. 10.21a identi-
cally. Hence we conclude that Eq. 10.27 represents a flow pattern
satisfying the conditions of continuity, momentum, energy, and irro-
tationality.
Streamlines. To determine the form of the streamlines, we make use
of the definition that the slope of the streamline is identical with the
slope of the velocity vector, so that, referring to Fig. 10.1, we write
The values of v/V, and u/V, are known in terms of x and y through Eqs.
10.28 and 10.30, but on substituting these values into Eq. 10.32 there
seems to be no simple way of separating variables and thus integrating
to find the equations of the streamlines. However, this difficulty may be
avoided by introducing approximations in keeping with the assumptions
of small perturbations. We write the identity
312 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 10
Now, substituting u / V , and v/VW from Eqs. 10.28 and 10.30 into
Ea. 10.33, and rearranging, we obtain
hI
dy = - 2 r (sin ?)
2rx
e - 2 T h l z dz - 277~
This may now be integrated directly to give the equation of the stream-
line,
y = constant +h (10.34)
Streamlines f
\
FIG. 10.2. Flow past wave-shaped wall, showing streamlines and pressure distri-
bution a t surface. Amplitude of wave in wall is esaggerated.
(a) Incompressible flow (M, = 0).
(b) Subsonic flow (M, < 1).
cam= 0 (10.39)
From Eq. 10.38 i t is evident that on the solid wall the pressure is a
maximum in the lowest points of the troughs and is a minimum a t the
highest points of the crests. The variation in pressure along the contom
is one-half wave length out of phase with the shape of the wall (Fig. 10.2).
By tracing the cross-sectional area of a stream tube formed by two
neighboring streamlines, i t will be observed that the pressure is a maxi-
mum at the largest cross-sectional area and a minimum a t the smallest
314 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 10
FIG.10.3. Effect of Mach Number on pressure disturbance for flow past wave-
shaped wall.
(a) Peak pressure coefficient at wall.
(b) Lateral decay of peak pressure coefficient.
pressibility effects. Thus, since the lift and moment acting on a profile
are found by suitable integration of the pressures acting over the sur-
face of the profile, we might expect from Eq. 10.38 that as a first approxi-
mation the lift and moment coefficients would be inversely proportional
to 8. i.e..
allows us to predict the details of the subsonic flow past a body a t sub-
sonic speeds if we know the details of an incompressible flow past an
affinely related body. Similarly, the transonic and hypersonic similar-
ity rules show how experimental data for a certain body a t a certain M,
can be made applicable to a related body at a different M,.
Transformation of Variables Leading to Laplace's Equation. Suppose
that a thin, two-dimensional airfoil is moving a t the constant speed
- Vw through an infinite gas medium which is at rest a t great distances
from the airfoil. To an observer traveling with the uniform velocity of
the airfoil, the airfoil is stationary, and the gas a t great distances from
the airfoil flows past the airfoil with the uniform, parallel velocity Vm
(Fig. 10.1).
The flow pattern is in first approximation governed by the linearized
differential equation of the perturbation velocity potential,
u = av/ax = o
aty = fa:
= av/ay =o
For a profile of given shape, that is, with a contour given by a certain
function f(x,, y,) = 0, the solution of Eq. 10.21a, subject to the boundary
Art. 10.5 GOTHERT'S RULE 317
conditions given by Eqs. 10.40 and'l0.41, is a t best not easy. However,
it will be noticed that Eq. 10.21a is not very different in form from the ex-
act equation for incompressible flow, the latter being found merely by set-
ting p equal to unity. Also, it may be recalled that one of the devices
used for solving unfamiliar differential equations is to seek a transforma-
tion of variables which reduces the differential equation to a familiar
form for which a solution is available. These remarks suggest that by a
suitable transformation it might be possible to reduce Eq. 10.21a to
the Laplace equation and thus open the possibility of employing the
well-developed store of theory and experiment for incompressible flow.
A generalized transformation which leads to the desired result is
found by defining new variables x', y' and cp1(x', y') as follows:
where A,, Xu, and X, are constants whose values are to be determined.
We now transform Eq. 10.21a into the x', y', cp' system of variables
through the following relations:
which means that the x', y', p' system of variables describes the motion
of an incompressible fluid if A, and A, are so chosen that X,/X, = p.
318 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 10
where x,' and y,' represent the coordinates of the transformed profile
as given by
x,' = X,x, (10.51a)
Substituting Eqs. 10.52 and 10.50 into the boundary condition of Eq.
10.41, we get after rearrangement
dys' 1
- = -- (10.53)
dx' hzX, V,
If we now set
X; = XzXp (10.54)
and if we also define V,' to be equal to V , (which may be done without
loss of generality), we obtain
Comparison of Eq. 10.55 with Eq. 10.41 shows that the boundary con-
ditions for the incompressible flow past a new profile defined by x,', y,'
are satisfied when V,' = V, and = XZX,.
Art. 10.5 GOTHERT'S RULE 319
which means that the slope of the profile in the compressible flow pat-
tern is larger by the factor 1/0 than the slope of the corresponding pro-
file in the related incompressible flow pattern. But if the slope of the
profile a t each point is greater by the factor I/& it is also true that the
camber ratio, angle of attack, and thickness ratio must all be greater for
the compressible airfoil by the factor I/& Thus, denoting the camber
ratio by y , the angle of attack by a and the thickness ratio by 6, we have
"I Incompressible
1 I /
Camber Line
But
and hence, wit;h the aid of Eqs. 10.47 and 10.54, we find that
At equal values of x,/l, Cp, and Cp,' are related by Eq. 10.58. Therefore
the ratio CI/C{ must be the same as the ratio Cp,/Cp,'. A similar
argument leads to the same conclusion for the moment coefficient, Cn.
Thus we may write
which means that if the two profiles are related affinely as described
previously, the lift and moment coefficients for the compressible flow
pattern will be greater than for the related incompressible flow pattern
by the factor 1/b2.
We might inquire whether these results have any practical significance.
There has been accumulated fairly extensive theoretical and experi-
mental information on the properties of classes of affinely related pro-
files in incompressible flow, with systematic variations in camber,
thickness ratio, and angle of attack. If it is desired t o find the lift coeffi-
cient of one of these profiles a t a finite Mach Number M,, one first finds
(either theoretically or experimentally) the lift coefficient in incom-
pressible flow of an affinely related profile whose camber, thickness
ratio, and angle of attack are all smaller than the corresponding values
for the original profile by the ratio P. Then, by multiplying this lift
coefficient by l/p2, one finds the desired lift coefficient for the compress-
ible flow.
This method of projecting experimental data for incompressible flow
is sometimes awkward, since it requires incompressible data for a large
range of thickness ratios. It would be more convenient in many cases
to know how Mach Number affects the performance of a profile of fixed
shape. The Prandtl-Glauert rule, which is discussed in the next article,
yields information of this type.
then the forces acting in the two flows are related through
These equations mean that for the same profile at the same angle of
attack the pressure coeficient, lift coeficient and moment coeficient are all
aflected by Mach Number in proportion to the factor 1 / 4 1 - b2.
From Eqs. 10.60 and 10.61 it also follows that for a given profile the
lift-curve slope is proportional to I/@:
Now let us combine these relations with Eqs. 10.57, 10.58, and 10.59.
We find that if the compressible profile (unprimed) is affinely related to
the incompressible profile (triple primed) in such a way that
then
Rule 11:
Incompy~ssibleI[ Incompr=2~sibleIII
Cp/Cp = I/P Cp/Cp = I
Eq. 10.22 for Cp is in this case written with p ~ / p , on the right-hand side.
The latter ratio is given by the isentropic relations of Chapter 4 as
Substituting this into Eq. 10.22, and solving for tAL2, we obtain
However, since the linearized theory ignores terms of greater than lirst
order in Cp, the foregoing expression may be expanded by the binomia!
theorem to give
k-1
M L2~ M , ~- Mm2)MmZcP - . +
Substituting this into Eq. 10.68, and solving for Cp, we finally obtain
." fin,
is associated with the appearance of shock waves near the profile which,
by virtue of adverse pressure gradients, produce flow separation. This
effect is very similar t o stalling a t high angles of attack.
Fig. 10.8 shows the effect of M, on the pressure coefficient a t 30%
chord on the upper surface of NACA 4412 operating with an angle of
attack of -2". Here again it is evident that the Prandtl-Glauert rule is
correct as regards order of mag-
nitude, but underestimates the
effects of compressibility.
From these typical results we
may conclude that the Prandtl-
Glauert rule predicts the effects of
M, with good accuracy for very
thin profiles a t low Mach Num-
bers. For thick profiles and high
Mach Numbers, the rule provides
a guide to orders of magnitude
but is not accurate enough for FIG.10.8. Pressure coefficient versus M,
design purposes. at 30% chord on upper surface of NACA
4412, with a = -2". Comparison of
For additional experimental measured data with Eqs. 10.66 and 10.69.
dataillustrating thevalidity of the The curve marked C,* denotes the local
sonic condition (after Stack, Lindsey, and
Prandtl-Glauert rule, the reader is Littell).
referred to Chapters 11 and 12.
Laitone's modification of the Prandtl-Glauert rule is seen from Fig.
10.8 to predict the variation of Cp with M, somewhat better than does
the Prandtl-Glauert rule. It should be noted here that no single simi-
larity rule can be correct for all profiles, and hence the different similar-
ity laws compare more or less favorably with experiment, depending
upon the profile shape, angle of attack, position of measurement, etc,
It is not amiss here to point out that the Karman-Tsien pressure
correction formula (Chapter 11) is generally in better agreement with
experiment than is the Prandtl-Glauert rule.
FIG. 10.9. Similarity Rules I, 11, and I11 for thin profile in wind tunnel.
Using Eqs. 10.42 and 10.45, we see that the corresponding boundary
condition in the incompressible flow is
at y' = f h h : -
A,
[
hy av1(x', Y')
ayl ]y'= +Ayh
= 0
absent. I t may be shown (see Problem 10.10) thzt the effect of com-
pressibility on the interference pressure coefficient a t a fixed point on
a profile of fixed shape in a tunnel of fixed size is given approximately by
Thus, using one of the many methods for obtaining incompressible flow
patterns, it is possible with this rule to construct approximately a series
of corresponding patterns for compressible flow.
One-Dimensional Mow. The method given here refers t o flows which
are essentially two-dimensional and serves no useful purpose for one-
dimensional flow because in the latter case there are simple and more
accurate methods a t hand. However, we may employ one-dimensional
considerations to verify for a limiting case the validity of the rule as
given above. Suppose that we have a long duct of uniform width fol-
lowed by a slight change in area, after which the duct is again constant
in width (Fig. 10.11). Far upstream and far downstream of the change,
IA
IV
I
lM
- A+ d ~ i
V + dV;
M+dMl
Using- the Euler equation, the pressure coefficient for small changes
may be written
dp p V dV dV
C p = c = - - -2-
2p'V iPV2 v
Now, employing the one-dimensional considerations previously referred
to, we find from Table 8.1 that for isentropic flow,
and, accordingly,
PROBLEMS
10.1. Consider subsonic, two-dimensional flow with small perturbations, and
assume that the Prandtl-Glauert rule is valid.
If t h e free-stream Mach Number a t which the local speed on a profile first
becomes sonic is 0.80, what is t h e maximum negative pressure coefficient on the
profile at very low Mach Numbers?
10.2. The local pressure coefficient is defined a s
Show that the value of the pressure coefficient corresponding to the critical
velocity on the surface of a n airfoil is given by
is a solution of the linearized equation of motion and that it represents the flow
in a two-dimensional passage formed by one wavy wall (amplitude h and wave
length I ) and one straight wall, the two walls being a t a lateral distance H from
each other.
Explore the flow pattern corresponding to this solution.
Compare the streamlines and wall-pressure distributions with the corre-
sponding quantities when H is infinite.
Compare the influence of the straight wall at a fixed value of H in incompress-
ible flow with that in subsonic flow.
Determine how the "interference velocity perturbation" varies with &.
10.5. I n deriving the similarity rules of this chapter it was assumed arbi-
trarily that V,' = V,. Demonstrate that this is not a restriction on the
results by starting anew with the more general form, V,' = XvV , and showing
that the similarity rules thus obtained are the same as those previously obtained.
10.6. I n estimating the effect of distance from a profile on the disturbance
produced by the profile in incompressible flow, the shape of the profile is not
important a t very large distances from the profile. For large distances, then,
we might consider the disturbance of a doublet to be representative of that of
any profile.
The perturbation potential for a doublet a t the origin which is superposed
on a parallel flow V , is
v,x
= x 2 + Y2
10.7. The NACA profiles in the 4-digit series are defined in the following
way. The first digit represents the maximum camber, in per cent of the chord;
the second digit the location of the point of maximum camber, in tenths of the
chord; and the last two digits the maximum thickness, in per cent of chord.
Arbitrary camber equations and thickness functions, together with the number
of the airfoil, serve to specify the shape of the profile.
It is proposed to w d i c t the performance of the NACA 4412 profile at a Mach
Number of 0.6 and a t a2 angle of attack of 4", starting with data taken in a
low-speed wind tunnel.
For each of the three similarity rules, stipulate the number of the profile
which should be tested at zero Mach Number, specify in each case the angle
of attack for the low-speed test, and explain how you would extrapolate the
data to Mach Number 0.6.
10.8. NACA Report No. 646 gives the following data on lift-curve slopes
for the NACA 4412 airfoil, where M is the Mach Number and a is the angle of
attack in degrees:
where 1 is the profile chord, 2h is the height of the tunnel, and Au is the "inter-
ference" velocity.
From this and the similarity rules demonstrate that the effect of compressi-
bility on the interference velocity for a profile of fixed shape in a tunnel of fixed
10.11. Test Gothert's rule for flow inside two-dimensional ducts against
results obtained from a one-dimensional analysis according to the type of
argument presented in Art. 10.9. Assume that two long sections of straight
duct having diierent cross-sectional areas are joined together. Defining
line. Along 1-2, for example, 8, for reasons of symmetry, is zero, but
the velocity decreases from its free-stream value to zero at the stagna-
tion point. Along 2-3 the speed increases, while the angle 8 first increases
A Streamline
2,
4
Streamline
..
(a) (b)
FIG.11.1. Corresponding streamlines in:
(a) Physical plane.
(b) Hodograph plane.
NOMENCLATURE
Now, noting that p and 1C. are functions of x and y, and making use of
Eqs. 11.1 and 11.3, we may write
dp = cpz d x+ cp, d y = V(cos e d x + sin 0 d y ) (1 1.4a)
d# = +, d x + $C.yd y = ( p / p 0 ) V ( - sin e d x + cos e d y ) (11.4b)
Elimination of Physical Coordinates. Considering these relations as
two simultaneous equations for d x and d y in terms of d p and d*, we may
by elimination solve for d x and d y , thus obtaining
cos e po sin 0
dx = -d p - --d* (11.5a)
v P v
sin 0
dy = -drp
v + --d*cosv B
Po
P
Now, V and e are functions of x and y. Since p and $ are also func-
tions of x and y, it follows that p and J. may be considered to be func-
tions of V and 0. Thus, we may write
dx = (6
cos 0 po
vv - --$v)
P
sin
v
e
d v + (-cosV e PO -
P
Sme
v
h )dB (ll.7a)
Now we multiply Eq. 11.9a by cos 0 and Eq. 11.9b by sin 8, and add,
thus getting
ve = ( P O / P ) ~ + V (11.10)
Similarly, we multiply Eq. 11.9a by sin 0 and Eq. 11.9b by cos 8, a.nd
subtract, thus getting
To eliminate cp from Eqs. 11.10 and 11.13, we set 9ev = cpvb Thus,
noting that p and M depend only on V , and using Eq. 11.12,
+
which is the desired relation giving as a function of V and 8, since c is a
function of V only. This differential equation is linear in the dependent
variable +.
Procedure for Obtaining Physical Flow Pattern from Hodograph
+
Solution. Assuming that a function = +(V, 6) has been found which
satisfies Eq. 11.14, the procedure for determining the flow pattern is as
follows:
(i) The derivatives +v and +o as functions of V and 8 are found by
+
direct differentiation of = +(V, 8).
(ii) Using Eqs. 11.10 and 11.13, cpe and cpv are found as functions
of V and 8.
(iii) Eqs. 11.7 are integrated to give x and y as functions of V and 8.
Since we started with a solution for $ in terms of V and 0, we thus have
+ in terms of x and y. That is, we have the physical streamlines. If
these physical streamlines represent a practical problem of interest, the
solution is useful-otherwise not. The great disadvantage of the hodo-
graph method may now clearly be seen, namely, that the problem of
flow past given boundaries is not susceptible to direct treatment by the
hodograph method.
FIG.11.3. Illustrates definition of tan- FIG. 11.4. Limiting velocity for tan-
gent gas. gent gas.
- dp 2 dp 2 2 -
= p - = p c - pm2cm2 = pO2co2 = constant (11.15)
d(llp) dp
Integrating, we get
Now, combining Eq. 11.17 with Eq. '11.15 in various ways, we find
Equations 11.15, 11.16, 11.17, and 11.18 give the relations between the
various fluid properties for the tangent gas, and replace the customary
isentropic relations. The accuracy of these formulas of course depends
on how close the fluid properties in the flow field are to the free-stream
properties.
Let us see what happens during an acceleration of this fictitious gas.
As V increases, Eq. 11.18b shows that p decreases. But Eq. 11.17 shows
that c increases! From Eq. 11.18a, on the other hand, M increases, but
can never exceed unity. Hence the approximate methods outlined
here are applicable only to subsonic flow.
Limitations of Tangent-Gas Approximation. Another way of examin-
ing the limitations of the tangent gas is to note that for sufficiently
large accelerations the pressure may become negative. We proceed to
determine when the latter condition is reached. Setting p = 0, Eq.
11.16 becomes
Pm - Pm
2
1
Cm Pm P
We introduce this relation into Eq. 11.20, and thus obtain, subject to
the tangent-gas approximation,
Since Vi and V are uniquely related through Eq. 11.22, Eq. 11.21 may
now be expressed in terms of Vi rather than V. Noting that, according
to Eq. 11.22,
7
d1 - M - d
dV dVi
and replacing V-derivatives in Eq. 11.21 by Vt-derivatives, we get
From the compressible solution, J. = $(V, 8), the profile shape in.the
physical plane may be found by integration of Eq. 11.7. The compress-
ible-flow profile thus found will, in general, be diierent from the original
incompressible-flow profile. However, we can a t least make them
similar by insuring that the compressible solution merges continuously
into the incompressible one as M, approaches zero. This is done by
integrating Eq. 11.22 with such limits that V + V i as M, 4 0.
Two distinct steps are involved in this method: first, finding the re-
lation between the corresponding velocities in the compressible and
incompressible flows; second, finding the relation between the corre-
sponding profile shapes. Let us proceed now to the determination of
the relation between the corresponding velocities.
Relation Between Corresponding Velocities. Combining Eq. 11.21
with Eqs. 11.18a and 11.18b, we have,
dV; p dV dV cn dV
log vi = - log co +7
co + v 2
+ log K
v
346 HODOGRAPH METHOD Ch. 11
Now, combining the latter relation with Eq. 11.24a written for free-
stream conditions, we have
Eqs. 11.26 and 11.27 taken together comprise the relation between the
corresponding compressible and incompressible velocities. We now
come to the final and most important step, namely, finding a relation
between the corresponding pressure coefficients.
Art. 11.4 KARMAN-TSIEN PRESSURE CORRECTION FORMULA 347
Solving for p,/p, and substituting in the formula for C,, we get
cpi= 1 - (V;/Vim)2
Therefore, Eq. 11.26 may be written as
Substituting this expression for V/Vm into the previous relation for
C,, and simplifying, we finally obtain the Karman-Tsien pressure-cor-
rection formula :
17
Lower Critical Mach Number. Where the line for a particular value
of Cpi intersects the line marked M = 1 (sonic line) in Fig. 11.5a1there
may be read on the horizontal scale the value of M, for which the sonic
velocity is reached locally a t the point having the particular low-speed
coefficient C,,. It is known from experience that soon after supersonic
speeds are reached locally, there are usually rapid increases in the drag
coefficient, and the corresponding value of M, is called the lower critical
Mach Number, M, ,. Fig. 11.5b, which is based on the Karman-Tsien
pressure-correction curves of Fig. 11.5a, shows M, ,,as a function of C,.
Given the peak negative pressure coefficient for a profile a t low speeds,
Fig. 11.5b shows approximately the free-stream Mach Number a t which
compressibility effects begin to have a serious adverse effect on drag.
Comparison with Prandtl-Glauert Rule. For very small values of the
pressure coefficient, that is, for very small perturbations from free-
stream conditions, Eq. 11.28 reduces to the Prandtl-Glauert rule,
c, = cp,/l/l--M,2.
When C, is negative, corresponding to regions of velocity higher than
free-stream velocity, the Karman-Tsien (K-T) formula indicates a
greater influence of compressibility than does the Prandtl-Glauert (P-G)
formula. As a result, the K-T rule leads to a value of the lower critical
Mach Number smaller than that computed by the P-G rule.
For positive values of C,, on the other hand, the I<-T rule predicts
a compressibility effect which is smaller than that predicted by the P-G
rule. This is not of great practical importance, however.
Since the effects of compressibility depend on the local pressure coeffi-
cient, the K-T rule does not allow the derivation of a simple relation
showing the effects of compressibility on the force coefficients of a given
profile. If low-speed pressure distributions for the profile are available,
however, these may be corrected to higher Mach Numbers by the K-T
formula, and then, by integration, the high-speed force coefficients may
be found.
Accuracy of Kannan-Tsien Pressure-Correction Formula. The
reliability of the Karman-Tsien pressure-correction formula may be
tested by (i) comparing it with experimental data, or (ii) comparing i t
350 HODOGRAFH METHOD Ch.11
with more accurate analytical results for simple cases where this is
possible.
Fig. 11.6, which is a typical comparison (2) between experimental data,
the K-T rule, and the P-G rule, indicates good agreement between the
measurements and the K-T rule, but shows the P-T rule t o underesti-
mate the effect of compressibility.
ing the K-T formula to profiles of fixed shape we thus produce two errors
in opposite directions, and the experimental evidence seems to indicate
that they nearly cancel each other.
It is also well known (5) that the conjugate complex velocity is then
given by
Since this relation gives fii as an analytic function oi(zi) of zi, it follows
that z; is an analytic function of fii. Then, from Eq. 11.29a, pi i+i +
must be expressible as an analytic function of fii:
By changing the signs of the imaginary parts of Eqs. 11.29 and 11.30,
we get -
qi - iGi = P(&) = G(wi) (11.31)
and
w; = ui +
ivi = Vieie = dF/di$ (11.32)
Now Eq. 11.23b tells us that an incompressible solution 4i(Vi, 8) = 0
yields a corresponding compressible solution +(V, 8) = 0 if we merely
replaced Vi by V in the former equation, provided that Vi and V are
connected by Eq. 11.24. Noting that 8 is constant in the transformation,
it follows that
9 + i+= G(fi); p - i+ = G(w)
It seams reasonable to suppose that the functions G(o) and G(w) may be
respectively approximated with good accuracy by the functions G(fii)
and G(wi), inasmuch as w and wi are nearly alike. Hence we write
In Eqs. 11.5 there appears the term Po/P, for which we now write with
the help of Eqs. ll.18b and 11.24b,
With these relations we now form, noting that (cos I9 + i sin 8) = e",
Now, making use of Eqs. 11.29 to 11.34, this may be brought into the
fo m
compressible flow past the circle of radius b in the {%-planeof Fig. 11.10b
through the conformal transformation
To prove this, note that, a t the surface of the circle of radius b, the l,-
coordinate is 3.2 = bei", and hence
The complex potential for the flow past the circle in the r,-plane is
doublet:
F(li) = vim (ri + z)
well known c5) and is given by the superposition of a uniform flow and a
Changing the sign of the imaginary part gives the conjugate complex
(11.4Oa)
potential,
p(fi) = Vim (11.40b)
Now, since z; and Ci are uniquely related, we may write, after taking
the derivatives of Eqs. 11.38 and 11.40b,
1 - -2
Ti
Carrying out the integration, we get
x' =
(b + 3cos- u
(b2 -
2
b ( l - b ) sina + 2
tan-'
2b sin a
b2 - 1 I
-
The compressible profile is therefore approximately elliptical if X is
small. To find the half-chord, x',,,,
thickness, y',,,,
we set u = 0; and to find the half-
we set a = ~ / 2 . By taking the ratio of the two, we
find, after rearrangement, the thickness ratio of the compressible profile:
"I
b(b2 - 1)
tan-'
b2 - 1
&
6 =
'- Lb2 +
b(b2-1)2
2(b2 1) +ln
b- 1
b 1
-1+ (11.42)
Analysis of Eq. 11.42 shows that 6 > ail which means that the trans-
formed compressible profile is slightly thicker than the original incom-
pressible profile. For example, a t M, with ai = 0.095, Eq. 11.42 yields
6/6i = 1.02. This seems to be generally the case for other examples
treated in the literature. It appears, therefore, that if profile distortion
is neglected in applying the K-T method, the effects of compressibility
are exaggerated, which is exactly opposite to the sense of the error owing
to the linear pressure-density relation.
Art. 11.7 MISCELLANEOUS EXAMPLES 357
FIG. 11.11. Theoretical effect of wind tunnel walls on lower critical Mach Number
for circle (after Suzuki).
Mmcr 5 / i
-6 -
slightly greater thickness ratio but a slightly
smaller angle of attack. The most interest-
e I
ing results are shown in Fig. 11.12, where
the lower critical Mach Number is shown as
a function of thickness ratio and angle of
0
o I attack. For thin profiles, e.g., for thickness
Thickness Ratio ratios of the order of 0.10, M, ,,is very sensi-
G 11.12. Theofetical tive to angle of attack. For thick profiles, e.g.,
effect of thickness ratio and
angle of attack on lower for thickness ratios of the order of 0.5, the
critical Mach Number for critical Mach Number is insensitive to thick-
ellipse (after Tamatiko and
Tamada). ness ratio and to angle of attack.
7fs ....,
.,.... .
,..
Me
(a) (b)
FIG.11.13. Theoretical coefficient of contraction for jet escaping from sharp-edged
slit (after Busemann).
Both these relations are plotted in Fig. 11.13b, where there is plotted
also the results of typical experiments ('4) on round sharp-edged orifices.
The theoretical results follow the
experimental results in a qualita-
tive way. Both indicate a sub-
stantial rise in discharge coefficient
owing t o compressibility effects.
Elbows for Accelerated Mows.
A method for designing passages
-
which simultaneously turn and ac- 11.14, turning channel
celerate the flow, Fig. 11.14, has (after Carrier).
360 HODOGRAPH METHOD Ch. 11
(a) Mm
(b)
FIG.11.15. Theoretical drag of flat plate placed normal to stream (after Chaplygin).
Determine the per cent error for M = 0.1, 0.5, and 0.8. Note from these results
that the Karman-Tsien pressure correction formula (ignoring coordinate correc-
tion) is correct u p to terms of the order of M2.
(b) Determine corresponding expressions for p/p, in terms of M and &.
11.3. It is often said that the tangent gas is equivalent t o a perfect gas
with k = - 1.
(a) Compare the various tangent-gas relations with the corresponding isen-
tropic relations for a perfect gas with k = -1.
(b) Compare the pressure-specific volume curve of the tangent gas with that
corresponding to isentropic expansion of a perfect gas having k = -1.
(c) Compare the various tangent-gas relations with approximate formulas,
found by developing the exact isentropic relations for small changes from f r e e
stream conditions.
(d) What conclusions do you reach?
362 HODOGRAPH METHOD Ch.11
11.4. (a) Demonstrate, for the tangent-gas approximation, that the stag-
nation pressure is related to the free-stream properties by the formula
For a gas like air, should this formula be used with a value for k of 1.4 or of
(- 1 Explain.
(b) Show that if a perfect gas changes state along the line of states defined
by the tangent gas, the entropy change is given by
Verify the argument that if the processes are without friction, the tangent-gas
formulas may be thought of as referring to the nonadiabatic flow of a perfect
gas. Show that if the velocity changes are such that the pressure is changing
in a direction away from the free-stream pressure, heat must be rejected from
the gas; whereas if the pressure is changing toward the free-stream pressure,
heat must be added to the gas.
11.5. Determine the isentropic relations between density, pressure, sound
velocity, speed, and Mach Number, for a tangent gas having the point of tan-
gency a t the stagnation point.
11.6. Demonstrate that in the Karman-Tsien method,
represents the superposition of a uniform, parallel flow and a source flow, and
hence yields the incompressible flow past an infinite half-body.
Using the Karman-Tsien method, determine the corresponding profile shapes
for incompressible flow and for flow a t M, = 0.5.
11.8. The incompressible flow past a circular cylinder of radius R has the
complex potential
pi +
i*i = V,,
NOMENCLATURE
c speed of sound tional, steady flow is not
CD drag coefficient possible
CL lift coefficient peak negative pressure co-
CP pressure coefficient efficient
CP.w pressure coefficient corre- ratio of specific heats
sponding to M = 1 Mach Number
C ) limiting peak negative free-stream Mach Number
pressure coefficient, for which sonic velocity
above which irrota- is first reached on profiie
3
Art. 12.2 THE RAYLEIGH-JANZEN METHOD
where
Having thus determined po(x, y), the right-hand side of Eq. 12.1 is
evaluated approximately on the basis of this zeroth approximation,
but retaining terms only up to the order of MW2. Let u$ call the right-
hand side, as so evaluated, Fo(x, y). The next step is to obtain the first
approximation to compressible flow, pl(x, y), by solving the linear
Poisson equation,
+
c ~ z z P,, = Fo(x1Y) (12.3)
avo
-=
avo
Uw and - - 0 a t x = y = a ,
ax
avo
- -- 0 a t solid boundary (n is normal to boundary)
an
av1
- -- 0 a t solid boundary
an
Note that the method gives the compressible and incompressible flows
about the same body. The iteration method converges, however, only
if the flow is everywhere subsonic. It is not, however, restricted to
thin bodies nor to flows without stagnation points.
Important improvements in the details of carrying out the method
were developed by Kaplan and Imai.(6) For complete details as to
Art. 12.2 THE RAYLEIGH-JANZEN METHOD 367
the techniques employed by various investigators for numerous ex-
amples, the reader is referred to Reference 1 and the complete bibliog-
raphy contained therein.
We now turn to some of the significant results which have been found
with the Rayleigh-Janzen method.
Flow Past Circular Cylinder. Fig. 12.1 shows the velocity distribu-
tion for flow past a circular cylinder a t M, = 0.4. Besides indicating a
comparison with incompressible flow, the various curves show the differ-
ences between the first, second, and
third approximations and indicate 2 5
the degree of convergence of the
method. For 0 < 45", the increase
in local surface velocity is greater
15
for incompressible flow than for x
compressible flow; whereas, for Urn ,
90" > 6 > 45O, the opposite effect
holds. o5
The lower critical Mach Number
I I I I I
M, ,, i.e., the free-stream Mach 0 20 40 60 30 100 120
Number a t which sonic velocity is" 8.. Degrees
-
first reached local'~in the flow, is F I ~ 12.1.
. Velocity distribution on cir-
found by the third approximation cular cylinder (after Imai).
to be 0.404. The Prandtl-Glauert
rule together with the incompressible solution yields M, ,,= 0.418,
and the Karman-Tsien rule correspondingly yields M, ,,= 0.40.
The maximum local velocity, which occurs a t the point of maximum
thickness, is related to the free-stream velocity and Mach Number as
follows:
FIG. 12.2. Lower critical Mach Number for circular cylinder in wind tunnel (after
Hasimoto).
flow past a circular cylinder, and indicates that caution must be used
in interpreting wind tunnel results.
Circular Cylinder in Two-Dimensional Free Jet. For a circular cyl-
inder in a free jet of limited lateral extent, as in an open wind tunnel,
Lamla gives the following results (see Fig. 12.3):
The maximum increase in local velocity is seen to be less when the free
stream is limited than when the latter is infinite, and hence the lower
critical Mach Number is increased as the jet becomes more limited in
width. Notice that this effect is exactly opposite to that in a closed
wind tunnel.
The maximum widening of the jet owing to the presence of the profile
is increased as the result of compressibility effects.
Mow Past Elliptic Cylinder. An elliptic cylinder may be thought of
as approximating a wing of large span or a compressor or turbine blade,
Art. 12.2 THE RAYLEIGH-JANZEN METHOD 369
1.2
1 .o
0.8
0.6
%
0.4
0.2
-0.2
0 20 40 60 80 100
m e g Thickness Ratio, 8
(4 (b)
FIQ.12.4. Flow past elliptic cylinder (after Kaplan).
(a) Effect of Mach Number on pressure distribution.
(b) h w e r critical Mach Number as function of thickness ratio and angle of
incidence.
The most significant results which have been worked out for this type
of profile (lo) are summarized in Fig. 12.5. The lower critical Mach
Number for symmetrical Joukowski profiles is shown in dependence on
thickness ratio and angle of attack in Fig. 12.5b. The curves are like
thoee of Fig. 12.4b1for much the same physical reasons.
1.0
0.8
3
0.6
z
8
2 0.4
I 0.2
0
0 0.1 0.2 0.3 0.4 0.5 O 0.2 Q4 0.6 0.8 1
.0
Thickness Rotio t/!. Thickness Rotio. t/P
(4 (b)
FIG.12.5. Flow past symmetrical Joukowski profile (after Tamatiko and Umemoto).
(a) S h p e of profile.
(b) Lower critical Mach Number as function of thickness ratio and angle of
incidence.
(c) Effect of M, on lift coefficient.
2
(12.7)
Eq. 12.9 and of the equations for a,a,etc., provide higher approxima-
tions to the flow of a compressible fluid and may be expected to hold
good for larger departures from a uniform, parallel flow.
To apply the method to the flow past boundaries of specified shape,
Eq. 12.8 is first solved for the function pl. In finding p l it is only
necessary to determine the incompressible solution for the chosen
boundaries and then to apply the Prandtl-Glauert rule. Now, having
determined pl(x, y), the right-hand side of Eq. 12.9 may be expressed
as a function of x and y. In order to determine the function pz it is
then necessary to solve the resulting linear Poisson equation. The
process is repeated for higher approximations. Naturally the boundary
conditions for each approximation need be satisfied for each approxima-
tion only to the same order of t which is involved in the velocity poten-
tial.
The detailed calculations for specific problems are too lengthy to be
given here, but we shall summarize some of the results of practical
significance which have been found with this method.
Flow Past a Bump and Past a Symmetrical Profile Without Stagnation
Points. Kaplan (11) has developed the method for uniform flow past a
symmetrical shape with cusps at both the leading and trailing edges
(thus insuring against stagnation points). The results are applicable
to the flow past the cusped profile of Fig. 12.6a, and also to the flow
(a) (b)
F I ~12.6.
. Flow past Kaplan bump (after Kaplan).
(a) Application to profile without stagnation points.
(b) Applicat,ion to bump on wall.
<
near the bump on the plane wall of Fig. 12.6b. The latter interpreta-
tion, for example, would be useful in estimating the effects of roughness
or waviness on a surface in subsonic flow.
Fig. 12.7a shows the velocity distribution a t the surface of a profile
of 10% thickness ratio, for values of M, from zero to 0.83. As found
previously, compressibility tends generally to augment the local de-
partures from free-stream velocity. For the 10yo profile, the value of
M, ,is 0.742, i.e., a t this free-stream Mach Number the velocity is
exactly sonic a t the point of maximum thickness. It is found that the
solution is regular up to M, of 0.83, beyond which the series diverges.
Beyond this limiting Mach Number, denoted by M, Ii,, and sometimes
called the upper critical Mach Number, no irrotational potential flow
for the given boundary is possible. Presumably, for M, > M,li, shock
Art. 12.3 THE PRANDTL-GLAUERT METHOD 373
waves would appear in the flow field. With the limiting value M, li,
for the 10% profile, the zone of supcrsonic flow is shown in Fig. 12.7b,
1.5
14
I. 3
-..
;I 2
_
3
m
1. 1
0
8
3
, 1.0
>
0 9
O.e 0
0.7 m 0.2 0.4
(a)
zx/e
0.6 0.8 I
Subsonic / Subsonic
/ \
/ I \
Region of \
1
Supersonic
F l o w Without
Shock f o r M,=0.83
\
\
,
I
I
-
\
urn .- -- -..
Mm
(b)
FIG. 12.7. Flow past cusped profile with 1070 thickness ratio (after Kaplan).
(a) Velocity distribution a t surface.
(b) Region of supersonic flow for M, = 0.83
and is seen to extend to a distance on each side of about seven times the
half-width. This example is significant in that it demonstrates that
there is no theoretical reason to expect shock waves in a potential flow
merely because the flow becomes supersonic locally.
-7 1 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 12
Thickness R a t i o
(b)
FIG.12.8. Flow past cusped profile (after Kaplan).
(a) Maximum negative pressure coefficient. Also shown are the sonic line and
the limiting line beyond which steady potential flow is impossible.
(b) U ~ p e and
r lower critical Mach Numbers.
Flow Past a Circular Arc Profile. The results given above refer to an
uncambered profile. To investigate the effects of camber, Kaplan (I2)
studied the flow past a circular arc profile of zero thickness (Fig. 12.9a)
a t zero angle of attack, with the circulation adjusted to satisfy the Kutta
condition of a stagnation point a t the trailing edge.
give the Kutta condition a t the trailing edge of the major axis. The
results of this study are summarized by the following formula relating
the lift coefficients about the same elliptical profile in compressible and
incompressible flow:
with experiment. The curves of Fig. 12.10b are in fairly good agreement
with experimenta'l results on profiles having the same thickness coeffi-
cients as the elliptical profiles.
t/P=O.12
H/4=3.6
W 1 : 30
NACA 0014
- 1.0
-0.8
-0.6
CD
-0.4
.O 2
0
0.2
0 10 20 30 40 5 0 6 0 7 0 8 0 9 0 IOU
Percent Chord oercent Chord
FIG. 12.11. (a) Flow past NACA 0012 profile in wind tunnel.
(b) Streamlines and lines of constant Mach Number, as worked out by relaxation
method for M, = 0.75 (after Emmons).
(c) Theoretical and ex~erimentalDressure distributions on profile for M, = 0.60
(after Emmons; Amick). '
(d) Theoretical and experimental pressure distributions on profile for M, = 0.75
(after Emmons; Amick).
Art. 12.5 SOME MEASURED EFFECTS OF COMPRESSIBILITY 379
-1.2
-0.8
-0.4
0
0.4
0.8
1.2
0 20 40 60 80 100
Chord, Percent C h o r d , Percenl
(b)
- 1.6 -1.6
- 1.2 -1.2
-0.8 -0.8
-0.4 -0.4
Cp 0 CP 0
0.4 0.4
0.8 0.8
1.2 1.2
0 20 40 60 80 I00 0 " 20 40 60 80 100
Chord , Percenr C h o r d , Percent
(c) (d)
F I ~12.12.
. Pressure distribution on NACA 4412 profile a t angle of attack of l052'
(after Stack, Lindsey, and Littell).
(a) M, = 0.19.
(b) M, = 0.60.
(c) M, = 0.66 (note shock a t 55% chord).
(d) M, = 0.73 (note shock at 67% chord).
Woke
.
Low M a .,
Wake
-
---__ D
H i g h Mm
F I ~12.13.
. (a) Flow past profile a t low Mach Number.
(b) Flow at high Mach Number, with boundary-layer separation owing to shock
wave.
Art. 12.5 SOME MEASURED EFFECTS O F COMPRESSIBILITY 381
FIG. 12.14. Schlieren photographs of flow past NACA 23015 profile a t angle of
attack of 3" (after Stack). Flow from left to right.
(a) M, = 0.40. Flow entirely subsonic.
(b) M, = 0.60. Small zone of supersonic flow a t 25% chord, with weak repeated
shocks.
(c) M, = 0.70. Large zone of supersonic flow, with lambda shock and boundary-
layer separation.
(d) M, = 0.80. Extensive shock pattern with so much boundary-layer separa-
tion that profile is stalled.
FIG. 12.15. Schlieren photograph of flow past Llustsng wing at M, = 0.78, with
angle of attack of 2" (after Hilton). Flow from right to left.
Boundary -Layer
Losses
\
Shock
.-
c "7
g 30 Losses Losses
"7 L Mw=0.910
oa 20
ap 0 10
, n
ing to the initial appearance locally of supersonic flow), the drag and
moment coefficients remain nearly
constant, while the lift coefficient
increases approximately according
to the Prandtl-Glauert rule. At a
Mach Number of about 0.7, as
compared with M, ,,of about 0.58,
there is a "force break,," i.e., the
drag coefficient suddenly increases,
the lift coefficient suddenly falls
+
off, and the moment coefficient
C, Measurements
decreases. An extensive summary -
of such results is given in Refer- o C , Measurements
ence 20.
Application of t h e P r a n d t l -
Glauert Rule to Drag. Karman (21)
has suggested a method, based on
the Prandtl-Glauert rule, for pre-
dicting the effect of M, on drag
coefficient in the range of Mach w
Numbers where shock waves are
not present. It is assumed that if Calculated DI I.Curv
the surface distributions of pres-
sure coefficient are the same in an
incompressible and in a compressi-
ble flow, then the local skin-fric-
tion coefficient will be alike a t FIG. 12.18. CL and CD versus M, for
NACA 4412 prome a t a = lo 52', show-
corresponding points, and hence the ing comparison of calculated drag with
drag coefficient will be the same measured drag (after von Karman).
384 TWO-DIMENSIONAL, SUBSONIC FLOW Ch. 12
for both profiles. The drag coefficient of a given profile a t M, and with
a certain value of CLM,is, therefore, taken to be equal to the drag coef-
ficient in incompressible flow of an affinely related profile operating a t
an angle of attack to give the same lift coefficient CLM-;the incompressi-
ble profile must have its thickness and camber larger by the factor
1/- than the corresponding quantities for the original profile.
A comparison of this method with experimental results for the NACA
4412 profile a t a = 1.88' (Fig. 12.18) shows, for this example a t least,
that the method works well up to the value of M, a t which the "force
break" or "compressibility burble" takes place.
.I2
I I
.I 0
.09
.08
.07
CL
.06
.05
.o 4
.03
.o2
O
-=
'00.4 0.5 0.6 0.7 0.8 0.9 1.0
Mm
Mm
(b) Fro. 12.20. (Explana.tion on page 387.) (4
Art. 12.6 THE STREAMLINE CURVATURE METHOD 387
b& .,
remain high is about three times as wide as the range of CL for which
is high.
GL
FIG.12.21. Force-divergence Mach Numbers for NACA 66-210 profile designed
for CL of 0.2 (NACA).
If the profile is thin the equipotential lines are nearly parallel to the
y-axis. Hence, we may, with little error, write dn Z dy, and the condi-
tion of irrotationality accordingly becomes
Substituting this equation for p/p,, together with Eq. 12.14 for R, into
Eq. 12.13, we obtain
This may be readily integrated and solved for (V,/Vm - 1) in the form
FIG.12.23. Velocity distribution (at M, = 0.75) for flow past 10%-thick profile of
Fig. 12.7b. Curve marked "exact" is taken from Fig. 12.7a (after Perl).
At any point in the flow the complete velocity components in the x-, y-,
+
z-directions are, respectively, (V, u), v, and w.
Art. 13.2 GOTHERT'S RULE 395
A review of Art. 10.2 will show that the most stringent assumptions
underlying Eq. 13.1 are that
( M m ) 2 ( ~ / V<<m1); Mm2(v/Vw)<< 1 ; M ~ ' ( W / V<<~1) (13.3)
Mm2(u/Vm)/(1- Mm2) << 1 (13.4)
Equation 13.4 bars from the present analysis the transonic region, that
is to say, free-stream Mach Num-
bers near unity. In the subsonic
region Eq. 13.3 shows that the
value of M , up to which the analy-
sis is accurate depends on the rela-
tive magnitudes of the perturba-
tion velocities produced by the
body.
Boundary Conditions. Let us
suppose that the fluid streams past
a thin body B (Fig. 13.1) whose
surface is defined implicitly by the
relation S ( x s , y,, z,) = 0. At in-
finity the boundary conditions are
where u,, v,, and w, are the perturbation velocities a t a point on the
surface S. On the surface S we may also write, however,
dS
as
= - dx,
as
+- as
dy, + -dz, = 0
axs ayS az,
Elimination of dx,,,dy,, and dz, with the help of the preceding relations
then gives us the boundary condition in the form
where u', Y', W' are the velocity components in the incompressible flow.
By applying the transformation formulas of Eqs. 13.7 and 13.9 to
the boundary condition of Eq. 13.5, we get
which means that in the incompressible flow, the flow is uniform and
parallel a t infinity.
Next, applying Eqs. 13.7 and 13.9 to the boundary condition a t the
body as expressed by Eq. 13.6, we obtain, after simplification,
Without loss of generality, we may set V,' = V,. Then Eq. 13.11 (by
Art. 13.2 GOTHERT'S RULE 397
comparison with the form of Eq. 13.6) is exactly the condition that the
incompressible flow follow the boundaries of B'. The latter, according
to Eq. 13.7, is thinner than B in all lateral dimensions by the factor 0.
Gothert's Rule. The bodies B' and B are affinely related, so that in
going from B to B', the thickness ratio in the x,y-plane, the aspect
ratio seen in the x,z-plane, the angle of attack seen in the x,y-plane, etc.,
are all altered in the same proportion, namely, they are reduced by the
factor p. Letting 6 denote the thickness ratio, and using the convention
that in affine transformations a change in 6 produces proportionate
changes in angle of attack, a , aspect ratio, B , etc., we may summarize
the entire analysis by the symbolic relation
or, in words, the pressure coe$cient at a given point for the compressible
flow at Mach Number M, past a body of thickness 6 i s 1/b2 times a s large
a s the pressure coeficient at the correspondzng point for incompressible
flow past a thinner afine body o j thickness P6.
This will be recognized as Rule I of Chapter 10. But, whereas for
two-dimensional incompressible flow the pressure coefficient for affine
thin profiles is approximately proportional to the thickness ratio, no
such simple rule applies to three-dimensional incompressible flow. For
three-dimensional flows, therefore, it is not possible to arrive a t the
simple Prandtl-Glauert rule showing the effect of Mach Number on the
pressure coefficient for a body of fixed shape.
In applying Eq. 13.13 to a compressible flow a t Mach Number t&,
it is necessary t o consider the incompressible flow past a different body
all of whose lateral dimensions are reduced by the ratio 4 1 - Mm2.
The pressure coefficient for the incompressible flow is found either from
systematic experiments or by means of analytical methods.
THREE-DIMENSIONAL, SUBSONIC FLOW Ch. 13
Aspect Ratio
of Revolution
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
1.o
Mm
(4
FIG.13.2. Flow past ellipsoids (after Hess and Gardner). (See explanation on page 399..
Art. 13.3 FLOW PAST ELLIPSOIDS 399
The lift and moment coefficients are found by integrating the pressure
forces acting on the surface of the body. Within the linear theory
Eq. 13.13 holds a t all corresponding points of the two flows. It follows,
therefore, that the rule expressed by Eq. 13.13 applies also to the lift
and moment coefficients.
Typical results of calculations of this type (2) are shown in Fig. 13.2.
ness ratio remaining fixed. It is clear that the pressure coefficient for a
given thickness ratio is very much less in three-dimensional flow than
in two-dimensional flow. This is the result of a three-dimensional relief
effect which arises because, in a three-dimensional flow, the stream may
deviate from the original direction in both the y- and z-directions,
whereas onljr the y-direction is allowable for a two-dimensional flow.
Thus a three-dimensional body produces a lesser disturbance in the
uniform, parallel flow, and, consequently, it has a smaller peak negative
pressure coefficient.
One of the very important conclusions to be drawn from this example
is that for practical problems the linearized theory is more reliable for
three-dimensional flow than for two-dimensional flow because the relief
effect makes the assumptions of Eq. 13.3 more likely to be met.
Influence of Aspect Ratio on Compressibility Effect. To bring out
more clearly the effect of M, a t various aspect ratios, the results of Fig.
13.2b are replotted in Fig. 1 3 . 2 ~to show the ratio of the peak negative
- pressure coefficients for compressible and incompressible flow past ellip-
soids of 15% thickness ratio. For infinite aspect ratio, corresponding to
two-dimensional flow past an elliptic cylinder, the curve of Fig. 1 3 . 2 ~
corresponds to the Prandtl-Glauert rule, namely, C, .-/I
For finite aspect ratios the effect of Mach Number is considerably less,
until for ellipsoids of revolution with small thickness ratios the pressure
coefficient is almost independent of Mach Number. For a thickness
ratio of 0.15 and an aspect ratio of 2 the effect of Mach Number is ap-
proximately midway between that of the ellipsoid of revolution and that
of the elliptical cylinder.
Lower Critical Mach Number. On Fig. 13.2b is plotted the curve of
u,,,/V, versus M, for which the value of u,, is exactly sonic. The
intersection of this curve with each of the aspect ratio curves gives the
value of M, , the lower critical Mach Number, at which sonic velocity
is first reached a t the surface of the ellipsoid. By plotting M, ,,versus
aspect ratio for constant thickness ratio, the curves of Fig. 13.2d are
obtained. This chart shows that the lower critical Mach Number is
markedly increased as the aspect ratio decreases to values less than
about 2.0. The increase in lower critical Mach Number is the result
of the three-dimensional relief effect, which reduces both the magnitude
of the incremental velocity and also the effect of Mach Number on this
magnitude.
Effect of Aspect Ratio on "Compressibility Burble." The lower critical
Mach Number, M, ,, is of practical significance because experience
has shown that, as the result of a shock wave-boundary layer inter-
action, the drag of a body rises when velocities on the surface substan-
Art. 13.4 BODIES OF REVOLUTION 401
tially exceed the sonic velocity. From the results on ellipsoids, there-
fore, it would be expected that the &TachNumber a t which the compressi-
bilitg burble appears would be increased as the aspect ratio is decreased.
No direct data for ellipsoids are available, but Fig. 13.3,(9)showing the
results of measurements on a wing having a rectangular planform and r
symmetrical pr'ofile of 12y0 thick-
ness ratio, gives evidence of the
beneficial influence of three-di-
mensional effects in delaying the
sharp drag rise associated with
compressibility. Corresponding
schlieren photographs are given
in Fig. 13.16.
It is also well known that shock
waves first appear on the wings
of high-speed subsonic aircraft
rather than on the fuselage or on
the tail surfaces. This is further
confirmation of the three-dimen-
sional relief effect.
.I
(13.14)
where L is the length of the body; x is the distance aft of the nose; r, is
the radius from the axis to the body surface; A = T T ~A'; = dA/dx;
A" = d2A/dx2;and so forth.
The results given by this approximate formula agree well with the
exact results for an ellipsoid of revolution, discussed later.
Disturbance Produced by Body of Revolution. To determine the
effects of wind tunnel wall interference, or to estimate the effects of a
body of revolution on the pressure coefficient a t large distances from the
body, we may assume that in incompressible flow the influence of the
body at remote points may be approximately represented by a three-
dimensional doublet. The incompressible velocity potential of the latter
is
x
(P-
(x2 + r2)%
so that, for a given body of revolution in incompressible flow,
u -
turbation velocity far from the body to be given approximately by
6'.
Now, referring to Fig. 13.4, let points 1 and 2 be corresponding points
in compressible and incompressible flow about corresponding bodies,
S and Sf, in the sense demanded by Gothert's rule; let point 3 refer to
incompressible flow past Sf, but a t the same physical coordinates as
point 1; and let point 4 refer to incompressible flow, a t the coordinates
of point 1, past the original body S. Then to find the effect of Mach
Number on the pressure coefficient for a $xed body a t a fixed point, we
first write the identity
Art. 13.4 BODIES OF REVOLUTION 403
A B c D
F I ~13.5.
. Flow past semi-infinite half-body of revolution (after van Driest).
(a) Nomenclature.
(b) Comparison of Gothert's rule with experimental data.
are solid lines indicating the effect of Mach Number on pressure co-
efficient as found by Gothert's rule. To apply the latter, the theoretical
solution for incompressible flow was first worked out as a function of
thickness ratio by the method of sources and sinks, and then Gothert's
rule applied in the usual manner.
The agreement between theory and experiment is on the whole fairly
satisfactory, even up to Mach Numbers of 0.9. Thanks to the several
effects mentioned earlier in reference to the flow past ellipsoids, the
method of small perturbations is seen to be considerably more reliable
Art. 13.4 BODIES OF REVOLUTION 405
In the absence of better information, Eqs; 13.17 and 13.18 may also
be used as a guide for estimating the peak negative pressure coefficient
on slender bodies of revolution of shapes other than ellipsoidal. For this
406 THREE-DIMENSIONAL, SUBSONIC FLOW Ch. 13
FIG.13.6. Substantially exact theoretical results for flow past ellipsoid of revolu-
tion (after Schmieden and Kawalki).
(a) and (b) Maximum perturbation velocity.
(c) Ratio of maximum perturbation velocity for compressible flow to that for
incompressible flow.
hypothesis that Eq. 13.18a may be used as a general rule for thin bodies
of revolution. I n Reference 20 it is also shown that, within the approxi-
mations of the perturbation theory,
- (-Cp,)mx = 2 1 (13.18b)
RESULTS OF EXACT THEORETICAL ANALYSIS. Substantially exact
results for potential, compressible flow past ellipsoids of revolution (3)
are shown in Fig. 13.6. For thickness ratios below 0.20, the approximate
Art. 13.4 BODIES OF REVOLUTION 407
formulas of Eqs. 13.17 and 13.18 are found to fit these results to within
a few per cent.
Figs. 13.6a, b show, as was mentioned previously, that the maximum
dis'turbance velocity produced by a thin body of revolution is small
compared with that of a two-
dimensional body. Fig. 1 3 . 6 ~
,
Drag. Fig. 13.9 shows the results of drag measurements (12) made on
a bomb.
This body of revolution has a comparatively large thickness ratio,
namely, 6 2 0.25. Hence i t is to be expected that the drag will be pre-
dominantly form drag, i.e., pressure drag. In the range of M, where
Mm
the flow is entirely subsonic, the pressure drag should, therefore, exhibit
a variation with M, somewhat similar to the variation of pressure coef-
ficient for inviscid flow. A comparison of the bomb drag in the range of
low Mach Numbers with the pressure-coefficient curve for 6 2 0.25 in
Fig. 13.6 shows that the rise of drag coefficient is about twice as large as
this argument would indicate, most likely because the shape of the bomb
produces higher incremental velocities than does an ellipsoid of like
thickness ratio.
According to Fig. 13.6, for 6 0.25, the lower critical Mach Number
is about 0.86. At about this point Fig. 13.9 indicates a very rapid in-
crease in drag coefficient, owing probably to the appearance of shock
waves.
.4rt. 13.5 SPHERES 409
13.5. Spheres
Gothert's Rule. The sphere is of some interest because of its simple
geometry and also because it represents the limiting case of a body of
revolution. Since a sphere is far from slender, Eq. 13.15 must be used
for the exact incompressible solution. Then, noting that 6 = 1 for a
sphere, Gothert's rule applied to Eq. 13.15 yields the prediction that for
compressible flow
0 30 60 90
0 , Deg.
FIG.13.10. Maximum perturbation ve- FIG. 13.11. Pressure dis-
locity for flow past sphere as function of tribution on sphere for com-
Mach Number. Comparison of exact so- pressible and incompressible
lution (after Kaplan) with Gothert's rule. flow (after Kaplan).
Since the lift and moment coefficients are found by integrating the
pressure coefficient over the entire surface of the wing, it follows from
Eq. 13.19 that the similarity rules for the lift and moment coefficients are
Lit-Curve Slope. This similarity relation, which holds for any plan-
form, compares wings of different aspect ratio and thickness ratio and
a t different angles of attack. To obtain a relation showing clearly the
Art. 13.6 WINGS OF FINITE SPAN 411
where the factor 0.9 takes account of an average departure for existing
wings from the theoretical value of 27r. Combining the last two relations,
we get
Al
We may now write the identity
The first ratio on the right-hand side is found from Gothert's rule; the
second is found from the approximate rule that the lift of affinely related
profiles is proportional to the angle of attack; and the third is found
from Eq. 13.21~. Thus we get
Fig. 13.13, which is based on Eq. 13.22, shows how the aspect ratio
modifies the effect of compressibility on the lift coefficient. For infinite
-
aspect ratio, Eq. 13.22 reduces to the Prandtl-Glauert rule for two-
dimensional flow, namely: for a fixed profile, d C ~ / d a I/-.
At the other extreme of zero aspect ratio, the lift-curve slope is inde-
pendent of Mach Number. For aspect ratios in the neighborhood of 3,
the augmentation of lift coefficient by compressibility is only about half
what it would be for infinite aspect ratio.
In a high-speed airplane, the tail surfaces have a much smaller aspect
ratio than the wing, and hence become relatively less effective a t high
412 THREE-DIMENSIONAL. SUBSONIC FLOW Ch. 13
FIG.13.13. Influence of aspect ratio on the way in which the lift-curve slope varies
with Mach Number.
which shows that the induced drag increases less rapidly with Mach
Number for wings of small aspect ratio than it does for wings of large
aspect ratio.
Experimental Results. The effect of Mach Number on the lift-curve
slope of a 12%-thickness profile with ~3 = 1.15 is shown in Fig. 13.14a.(15)
I t is seen that the Prandtl-Glauert correction factor for two-dimensional
flow is much too large and that the approximate theory developed in
this chapter agrees rather well with the measurements up to the point
where the lift-curve slope drops off as the result of local shock waves.
The moment-curve slope for the same wing is shown in Fig. 13.14b.
The change in the sign of the moment-curve slope a t M, E 0.75 has a
I
crucial bearing on aircraft stability a t high speeds.
Fig. 13.15 represents the results of measurements on wings having a
12%-thickness profile and various aspect ratios.(g) At low Mach Num-
Art. 13.6 WINGS O F F I N I T E SPAN 413
Mm Mm
(a) @)
F I ~ 13.14.
.
0012 profi le (after Gothert).
-
Ex~erimentalresults for wine havine- asuect
. ratio of 1.15 and NACA
ift-curve slope a t zero incidence versus Mach Number, showfing comparison
ith Gothert's ~ l e .
L
(4
NACA 0012 profile in wings of vari
414 THREE-DIMENSIONAL, SUBSONIC FLOW Ch. 13
FIG.13.16. Schlieren photographs of flow past wings with NACA 0012 profile at
4" angle of attack (after Lindsey and Humphreys). See also Fig. 13.3.
(a) a = m; M, = 0.78.
(h) = m; M, = 0.87.
(c) A( = 2; M, = 0.78.
(d) A( = 2; M, = 0.86.
the relative positions of the lift-drag polars because the relation between
lift and induced drag, Eq. 13.23, is substantially independent of Mach
Number. At higher Mach Numbers, and especially a t high lift coef-
ficients, local compressibility effects give rise to shock waves, but this
effect is smaller for wings of low aspect ratio because of the three-dimen-
sional "relief effect." Thus, a t M, = 0.85, Fig. 13.15d1 the order of
Art. 13.7 SWEPTBACK WINGS 415
observer to move along the span of the wing a t the uniform speed
Vmsin u,. The approach flow then appears t o be normal to the leading
edge, and has the speed V mcos a,. The tangential velocity of the original
flow, Vmsin u,, does not influence the pressure distribution on the wing,
but is important only in the determination of the frictional stresses
on the surface. Since only the normal component of velocity is significant
416 THREE-DIMENSIONAL, SUBSONIC FLOW Ch. 13
in the potential flow, it follows that the effective Mach Number of the
flow is M, cos cr,. Thus, although the free-stream Mach Number M,
may be of the order of unity or even greater, the effective Mach Number,
M, cos v,, may, through sufficient sweepback, be made small enough
to avoid the adverse effects of shocks.
In the coordinate transformation mentioned above, the effective thick-
ness normal to the leading edge is also increased, which is unfavorable
as regards the maximum local velocity. This effect, however, is con-
siderably smaller than that of reducing the effective Mach Number
normal to the leading edge.
On the other side of the picture, sweepback has two disadvantages:
First, the lift coefficient is lessened by reducing the normal component
of velocity, thus leading to larger wing areas. Second, severe structural
problems are associated with sweptback wings.
Sweepback may also be used advantageously to improve the efficiency
of propeller tips a t high speeds and to ameliorate compressibility prob-
lems with compressors pumping gases of very high molecular weight
(i.e., low sound velocity).
Another way of picturing the physical effects of sweepback is to im-
agine a wing in a rectangular wind tunnel. If there is no sweepback,
the wing produces a blocking effect in complete planes normal to the flow.
If the wing is swept back, however, there is a comparatively small block-
ing effect in any plane normal to the flow. Hence, a three-dimensional
"reljef" effect is present, inasmuch as the air may avoid the wing by
flowing laterally in two dimensions. The corresponding reduction in the
perturbation velocity has a cumulative effect in compressible flow be-
cause the effects of Mach Number are decreased as the incompressible
disturbance velocity itself is decreased.
Gothert's Rule. I n applying Gothert's rule of similarity, we compare
the. original compressible flow, Fig. 13.17a, with an incompressible flow,
Fig. 13.17bJ past a body whose longitudinal dimensions are unchanged,
but whose lateral dimensions are all decreased by the ratio 8.
Assuming infinite aspect ratio, the geometric relations between the
two wings are
1
tan a< = = - tan U, (13.26)
edge edge edge
or
($)m max
=
ti
(~)icosa; = - 1
+
li 4 1 tan 2 ai
(13.28)
Then, by substituting Eqs. 13.25, 13.26, and 13.27 into Eq. 13.28,
simplifying, and dropping the subscript c for the compressible flow, we
get
tll -
- tll
(13.29)
+tan 1 - M,' +
tan 2 a
For the same profile in incompressible flow (b = I), we get
a x - (-Cp)
- max -
-
41 + tan a
2
(13.31)
m a ( C dl - Mm2 + tan 2
a
F I ~13.18.
. Effect of sweepback on the way in which the maximum perturbation
velocity for thin elliptical profiles varies with Mach Number.
If the wing is "sheared back" in such a way that the profile parallel
to V , remains constant, then Eq. 13.29 rather than Eq. 13.32 is to be
used. Since shear-back increases the thickness normal to the leading
edge, this type of sweepback is not as effective as bend-back. As an
example, for an elliptical profile of 10% thickness ratio parallel to V,,
attack of -2". Data are shown for three conditions, with different
values of u and M, but with the same value of M, cos u. The ordinate
represents the pressure coefficient based, in each case, on the corre-
sponding normal component of velocity. The principle under discussion
is, a t least approximately, verified by the fact that the three sets of
data are nearly superposed when plotted in this manner.
Fig. 13.21a, from the same set of experiments, demonstrates the same
principle by showing that curves of C~/cos' u versus M,cos u are almost
-
1 ~CLN
(da)
2lr
cos .
where CLN is the section lift coefficient of the profile, measured with
respect to the speed and chord in a plane normal to the mean chord line;
422 THREE-DIMENSIONAL, SUBSONIC FLOW Ch. 13
FIG. 13.23. Correlation of lift-curve slopes for various planforms (after Diederich).
and a is the angle of attack (in radians) of the entire wing measured in a
plane parallel to the axis of symmetry of the wing.
Using CL to denote the lift coefficient of the entire wing, based on V,,
the total planform area, and the incidence a, the semi-empirical law
propos
- F
- - (13.33)
( d C ~ ~ / d cos
a) u
- 2- +, F-
4
1;
y - ' $
A comparison of this formula with accurate theoretical and exper-
imental results is shown in Fig. 13.23 for a wide variety of planforms.
The validity of Eq. 13.33 as a general correlation rule is seen to be
quite satisfactory.
PROBLEMS
13.6. It has been shown by Bollay [Z.a.M.M., Bd. 19 (1939), p. 211, that
for incompressible flow past a flat plate of vanishingly small aspect ratio, the
lift coefficient is related to the angle of attack by the formula
Using Gothert's rule, find a similar relation for compressible flow. Compare
your results with the discussion in Art. 13.6.
13.7. Consider a wing of infinite span with 45"-sweepback and a 10%
elliptical profile. Assuming that the tangential component of velocity is of no
consequence, find the value of M, for which the local sonic velocity is first reached
at the surface of the wing. Compare your result with Fig. 13.19.
PART V
SUPERSONIC FLOW
NOMENCLATURE FOR CHAPTER 14
angle of attack Cartesian coordinates
speed of sound
drag coefficient Mach angle
lift coefficient thickness ratio; angle
moment coefficient coordinate normal to chord
pressure coefficient turning angle of streamline
drag (radians)
amplitude of wave-shaped wall 180 e / ~
ratio of specific heats coordinate along chord
wave length of wave-shaped mass density
wall; also chord of airfoil surface slope of profile with
lift respect to chord line
Mach Number perturbation velocity potential
counter-clockwise moment signifies free-stream condition
about leading edge signifies upper surface of airfoil
pressure signifies lower surface of airfoil
pressure index (Eq. 14.20)
surface area of profile signifies step along Mach wave
area between chord line and of family I, or across Mach
profile surface wave of family II.
perturbation velocity compo- signifies step along Mach wave
nents of family 11, or across Mach
velocity wave of family I.
Chapter 14
and note that au/dx denotes longitudinal accelerations while a v / & ~is
connected with the convergence or divergence of the elementary stream
tube formed between two neighboring streamlines, we see that this
equation represents the already familiar fact that an acceleration re-
quires an increase in cross-sectional area for supersonic flow and a de-
-
crease for subsonic flow. Because av/ay and au/ax are of the same sign
for supersonic flow and of different sign for subsonic flow, the resulting
flow patterns are of different type. As in the one-dimensional simplifica-
tion, the long-familiar concepts of incompressible flow prove to be mis-
leading if applied to supersonic flows.
For subsonic flow Eq. 14.1 is said to be elliptic in type, and in fact
may - be reduced to the Laplace equation by the transformation given in
Chapter 10. Hence subsonic flows are, a t least qualitatively, similar to
incompressible flow. For supersonic flow Eq. 14.1 is said to be hyper-
bolic in type, and may be reduced to the classical wave kquation by a
simple transformation. Hence, linearized supersonic flows have many
properties akin to the properties of vibrating metal bars, air columns,
and strings. A fuller explanation of the mathematical properties of
elliptic and hyperbolic differential equations is presented in Appendix A.
Solution to the Differential Equation. We shall now show that for
supersonic flow the general solution to Eq. 14.1 is given by
c p = c p l ( x + m y ) + cp2(x- a?/ (14.4)
where cpl is an arbitrary function of the argument (z 1- +
y),
and cp2 is an arbitrary function of the argument (x - m y ) .
To demonstrate that this is indeed the solution, we first examine the
particular solution cp,. Taking derivatives,* we have
Then
u = ap/ax = v,(flf+ jzf)
a~ i
= jlldx +m j l f dy, etc.,
Within the assumptions of the linear theory, this may also be written a3
x- d ~- 1,y =~constant
These form a family of parallel straight lines with slope dyldx =
1 Referring to the velocity triangle of Fig. 14.1, we see
F I ~14.2.
. Flow with only left-running Mach waves present.
F I ~14.3.
. Flow with only right-running Mach waves present.
(Fig. 14.2b) were made solid. Then ABCD might be considered the
surface of a body with unbounded flow above the surface. We now
inquire which of the two solutions (jl or f2) is valid, or whether some
combination of the two might be valid. The answer depends partly on
the boundary conditions. If we specify no upper boundary to the flow
and that the flow far upstream is uniform and parallel, then the function
fi must be everywhere constant, for otherwise the flow far upstream
.
could not be uniform and parallel. From another point of view, if we
consider the Mach lines as representing the lines of propagation of
pressure waves produced by changes in direction at the solid boundary,
these waves can physically be propagated only downstream in super-
sonic flow, and hence only the pressure waves corresponding to the
function f2 can appear in the solution. Accordingly, the flow pattern
must be as indicated in Fig. 14.2b.
By similar arguments, if the region above ABCD were made solid; the
flow pattern would be as shown in Fig. 14.3b.
434 SUPERSONIC FLOW WITH SMALL PERTURBATIONS Ch. 14
If the curve BC represented a thin airfoil moving a t supersonic speed
through still air (or, equivalently, located in a supersonic wind tunnel
having a uniform, parallel air stream), the flow pattern would be as
shown in Fig. 14.4. The pressure distribution on the upper and lower
FIG.14.4. Flow past thin cambered profile. Above the profile the flow is affected
only by left-running Mach waves; below, only by right-running Mach waves.
surfaces of the airfoil (at each point of which the slope of the stream-
line is known) may easily be calculated from Eq. 14.11. That the flow
downstream of the airfoil must have the direction and pressure of the
upstream flow follows from the fact the two streamlines which join a t
the trailing edge must have the same pressure and direction; from the
two forms of Eq. 14.11, this will be so only when the flow is as shown
in Fig. 14.4.
the left-running Mach lines. Now, since (pl depends only on the argu-
+
ment (z ' d ~-, 1 ~y), the lines of constant (pl are all inclined a t the
angle -a, to the flow;similarly, the lines of constant (p2 are all inclined
a t the angle or, to the flow. We may then consider the flow field to be
covered with these lines, as shown
in Fig. 14.5; the specific nature of
the flow pattern then depends on
(x -= y = const.)
Lines of Constant rp,
the numerical magnitudes of (pl n n n n n n
and 9, attached to each of these
lines.
Suppose that in a small region
only the function (p2 undergoes any x
FIG. 14.6. Velocity vector diagram for a streamline influenced by left-running pres-
sure waves.
It may be seen by comparison that these relations agree with Eq. 14.11.
Four Basic Waves. I n constructing solutions by means of waves, we
may synthesize a flow pattern corresponding to specified boundary
conditions out of four fundamental types of changes, depending (i) on
whether the wave is of family I or of family 11, and (ii) on whether the
stream is accelerated or decelerated. These four types of changes are
illustrated in Fig. 14.7. From these sketches and Eqs. 14.12 and 14.13,
we see that the pressure increases for a deceleration and decreases for an
acceleration, as required by momentum considerations. Furthermore,
during a compression the streamline bends tmard the Mach line,
whereas during an expansion the streamline bends away from the Mach
line; by considering the cross-sectional area of a stream tube formed by
two neighboring streamlines, it will be seen that in a compression the
area decreases while in an expansion the area increases, as might be
expected from simple one-dimensional considerations for supersonic flow.
When a streamline is influenced by a wave, i t may be seen from
Fig. 14.6 that the velocity component parallel to the wave is unaltered.
Only the velocity component normal to the wave is changed in magni-
"i .\.
Acceleration
\
./'
-A
\
Deceleration
\ -u
Acceleration Deceleration
\ -ne
/-
F I ~14.7.
. Four basic types of Mach waves.
tude. Or, stated differently, the vector change in velocity is in the direc-
tion for which the component of velocity is equal to the local sound
velocity.
Fig. 14.8 shows the compression Mach waves of family I and family
I 1 generated a t the nose of a thin wedge-shaped body.
FIG.14.8. Shadow pliotograrh oI flow (from left to right) past thin wedge-shaped
body (M.I.T. Gas Turbine Laboratory).
438 SUPERSONIC FLOW WITH SMALL PERTURBATIONS Ch. 14
r;)
the streamline. For the solid wall, as an example,
h 2nx,
= = - 2~ - sin -
S 1 I
and, using Eq. 14.12b, we get
cach of which there are simple-wave patterns. From Eqs. 14.12 and
14.13, therefore, the local pressure at any point on the surface depends
only on the local inclination of the tangent to the surface.
Pressure Distribution. Let C;, 7 be a system of coordinates which lie
in the airfoil and which are therefore independent of angle of attack,
and for convenience let [ lie along the chord line. Also, let a denote the
local inclination of the surface measured from the chord line, and let
subscripts U and L refer to the upper and lower surfaces respectively.
Note that 7~ is measured upward and 7~ downward; similarly, uu is
measured counterclockwise and aL clockwise.
Referring to the geometry of Fig. 14.10b, the local directions of the
streamlines a t the surface, relative to the undisturbed flow, are
From Eqs. 14.12 and 14.13 we find the local surface pressures to be
Subtracting the second from the first of this pair of equations, we get,
for the pressure difference at any chordwise location,
inasmuch as
and, likewise,
ituL
s lzuu
dl 0; d [ s O
- -
where UL' and uu2 are the mean squares of the slopes, as defined by
and depends only on the angle of attack. Since the angle of attack is,
by Eq. 14.14a, directly associated with the generation of lift, this part
of the drag is known as the wave drag due to lijt, or, following the nomen-
clature of subsonic fluid dynamics, the induced drag:
and
where SL is the area between the lower surface and the chord (positive
when the lower surface lies below the chord), and Su is the area above
the chord. With these integrals we obtain
and
Center of Pressure. Suppose that l,,,, denotes the distance aft of the
leading edge through which the resultant force is applied. To first-
order effects,
c . . - 32
3n = LI,.,.; or - - -
1 Ll
dCdda 4 / d M W 2- 1 -/4
CD; 4 a 2 / d M W 2- 1 4 a 2 / d M m 2- 1
--
cot, 0 4 6 2 / d ~ , 2- 1
af(for CD, = 0) 0 6
(LID),,, (for C D , = 0) m 1/26
--
dCm/da 2 / d M W 2- 1
L.,,./l 0.5
O
'w
the results obtained thus far and draw some general conclusions.
LIFT. From Eq. 14.14b we see
that the lift-curve slope, d C ~ / d a ,
is the same for all thin profiles a t , c,
- V)
Prandtl-Glauert rule and on the well known value of 2~ for the theo-
retical lift-curve slope of a flat plate in incompressible flow.
DRAG. It may be seen from Eq. 14.15~that at small angles of attack
the thickness and skin-friction drag predominate, whereas a t large angles
of attack the induced drag predominates.
For a given profile a t a fixed Mach Number, Eqs. 14.14a and 14.15~
define a polar relation between C;, and CD with the angle of attack as a
parameter. As in subsonic flow, the lift-drag polar has the form of a
\
\
y\
\
\
6 \q\. ,
-
CL
CD
4
10% Double Wedge
4 6 8 I0
o. Degrees
F I ~14.13.
. Lift-drag polar for typical FIG.14.14. &ft-drag ratios for flat-
supersonic profile. plate and double-wedge profiles, show-
ing effect of skin friction.
parabola. Fig. 14.13 shows the lift-drag polar for a symmetrical double-
wedge profile of 10% thickness ratio operating a t M, = fi. Fig. 14.14
shows how the lift-drag ratio of the 10% double-wedge and of the flat
plate vary with angle of attack a t M, = 42. The angle of attack for
maximum lift-drag ratio, together with the value of (LID),,, are inde-
pendent of Mach Number and depend only on the shape of the profile,
provided of course that CD,= 0. It is clear from Eq. 14.15~that, from
the aerodynamic point of view, i t is desirable to use profiles with small
thickness ratios in order to minimize the thickness drag.
CENTER OF PRESSURE. The center of pressure is in the neighborhood
of the mid-chord point for profiles symmetrical about the chord, whereas
for subsonic flow the center of pressure is a t about 25% of the chord aft
of the leading edge. If an aircraft must fly a t both subsonic and super-
sonic speeps, therefore, the control problem will be seriously complicated
by the large changes in aerodynamic moments and in hinge moments of
the control surfaces as the aircraft passes from subsonic to supersonic
Art. 14.6 SUPERSONIC AIRFOILS 447
speeds. We also conclude from Eq. 14.19 that the control problem is
simplified if the profile is symmetrical about the chord line, for then the
center of pressure does not change with angle of attack. To first-order
effects the center of pressure is independent of Mach Number.
The foregoing generalizations, it should be emphasized, rest on the
linearized theory. Consequently they are not entirely correct. More
accurate methods of calculation are presented in Chapters 15 and 16.
For the present it will suffice to state that the linear theory is reasonably
correct as regards the lift and drag, but is unreliable as regards the
center of pressure.
Comparison with Experiment. Fig. 14.15 shows the theoretical and
experimental (1) surface-pressure distributions on a biconvex circular-arc
profile of 10% thickness ratio. The agreement is quite good for small
deviations from the free-stream direction and is very poor for large
deviations in direction, as might be expected from the assumptions of
the linear theory. Schlieren photographs indicate that the curious
variation of C, near the trailing edge of the lower surface is the result
of boundary-layer interaction with the oblique shock wave which,
according to a more exact theory, must be present near the trailing
edge (Chapter 16).
Although the pressure distributions are in poor agreement with the
linear theory, the errors are largely self-canceling as far as the resultant
forces are concerned, since the respective areas between the experimental
and theoretical curves for the upper and lower surfaces are nearly the
same. For this reason, the agreement as to lift and drag coefficients
is much better than the pressure distribution would indicate. For
\ example, (Figs. 14.16a and 14.16b show that the linear theory furnishes
a good approximation to the lift and drag coefficients. However, the
448 SUPERSONIC FLOW WITH SMALL PERTURBATIONS Ch. 14
- 8.6.3 %
Mm=2,.13
0 ~ e r r i ' sData
- Linear Theory
o, Degrees
To sum up, the linearized theory is useful for showing in a general way
the properties of supersonic airfoils. In addition, it furnishes a simple
method for approximating the lift and drag coefficients with fair accu-
racy.
Flat-Plate Airfoil in Subsonic and Supersonic Flow. I t is instructive
to compare a flat-plate airfoil in a frictionless fluid for incompressible
flow with the corresponding situation for supersonic flow. The stream-
line patterns are shown in Fig. 14.17. In supersonic flow the static
(0 )
/
/
/ - /
/ (b)
\ / -
b
\
b \. --
\ \
F I ~14.17.
. Streamlines for flat-plate airfoil.
(a) Subsonic flow.
(b) Supersonic flow.
sonic flow, but none in supersonic flow. Near the nose of the flat plate
in incompressible flow the velocities approach infinity and there are
infinite LLsuction
pressures," but no singularities of this sort appear in the
supersonic pattern. In incompressible flow the resultant force on the
plate is normal to the direction of the distant flow, that is, there is no
drag (the dlAlembert Paradox) ; in supersonic flow the resultant force is
normal to the plate, that is, there is a wave drag even though the fluid is
frictionless. The center of pressure is a t about 25% of chord for incom-
pressible flow, but a t about 50% of chord for supersonic flow. These
comparisons again illustrate the complete diierence in character be-
tween subsonic and supersonic flow.
The local pressure index P of course determines the local Mach Num-
ber through the isentropic relationships.
Art. 14.7 REFLECTION AND INTERSECTION OF WAVES 451
FIG. 14.20. Schlieren photographs of wave reflection (M.I.T. Gas Turbine Labora-
tory).
(a) Compression waves generated by wedge are reflected in like sense a t duct
walls. Reflected waves are subsequently reflected in like sense a t wedge
surface.
(b) Compression wave generated by wedge is incident on a corner, the turn-
ing angle of which is slightly greater than that of the wedge. The compres-
sion wave is more than canceled by the expansion generated a t the corner,
and the net result is a very weak expansion emanating from the corner.
If we first add and then subtract these two equations we obtain the
desired answers in the form
( b ) Underexponsion
that the wave and flow patterns undergo cyclic changes which extend
to infinity in the absence of friction. In adjusting itself to the back
pressure, the jet in a sense "overshoot~s,"for there are zones in the jet
where the pressure is higher than the back pressure.
Fig. 14.2313 shows the wave pattern and flow properties for an under-
expanded jet, i.e., for the case where the exhaust-region pressure is less
then the exit-plane pressure.
Example of Supersonic Flat-Plate Biplane. Fig. 14.24 shows h,ow the
linearized theory may be used to determine approximately the aero-
dynamic properties of a supersonic biplane. The particular example of
(a) (b)
F I ~14.24.
. Flat-plate biplane a t 1" incidence.
(a) Wave pattern, streamlines, and pressure distribution. Dashed lines show
expansion waves; full lines compression waves.
(b) Ratio of lift coefficient of biplane (based on area of both plates) to that of
monoplane.
the sketch is for an angle of attack of one degree. From the pressure
distribution of Fig. 14.24a, it may be seen that over certain areas of the
plate there is no net force acting, whereas over the remaining areas the
force per unit area is the same as for a single flat-plate airfoil. From the
A
geometry of the wave pattern, it is found that the fraction of the area
over which there are pressure differences depends on the parameter
d m -21 Fig. 14.24b indicates how the ratio of lift coefficient
for the biplane to lift coefficient for the monoplane a t the same angle of
attack varies with the parameter t d M m 2- 1/1. Further consideration
shows that the location of the center of pressure also depends in marked
degree on the parameter t d M m 2- 1/1. From the viewpoints considered
here, the flat-plate biplane is inferior to the monoplane, but for thick
profiles the biplane has certain advantages over the monoplane.
The method of applying the linearized theory as given in this article
should be considered as a simple and rapid means for obtaining qualita-
tive results and rough quantitative results. Accurate numerical results
may be obtained by this method only when there are very small pertur-
bations from a mean paraJ.Ie1 flow. For .large changes in velocity and
456 SUPERSONIC FLOW WITH SMALL PERTURBATIONS Ch. 14
PROBLEMS
14.1. Consider a nearly uniform, parallel flow with the following perturba-
tion velocity potential:
p=O for (Z - ~ M , -
Z 1 v) < 0
where A is a constant.
(a) Investigate the streamlines and pressure distribution.
(b) Show that when A approaches zero, the flow pattern is that corresponding
to flow from left to,right above a wall which has a sharp corner with a turning
angle c.
14.2. If the value of pz - p, across a weak pressure wave is calculated from
the linear theory, and if the values of Mz and of p,/pl are calculated from the
isentropic relations, it is found that the equation of continuity is violated across
p~~~~~
the wave. Explain why this is so.
,I
foil
airfoil
is shown
symmetrical
in the sketch.
about the
Thechordair-
line.
PROB.14.3. Find expressions for the lift, drag,
and moment coefficients in terms of
the Mach Number, the constants a1 and a2, and the thickness ratio 6 = t / l .
Show that for a fixed thickness ratio, the wave drag due t o thickness is a
minimum when a1 = a2 = 0.5.
Art. 14.7 PROBLEMS 457
14.4. Consider a thin, curved profile of parabolic shape (see sketch), ex-
pressed by the curve, y = -h(x/t)'. The leading edge of the profile is tangent
to the direction of the oncoming air stream. Using the linearized theory,
(a) Find expressions for the lift and drag coefficients, for the lift-drag ratio,
and for the moment coefficient. all in
~ ~
14.6. A certain tm-o-dimensional passage has one straight wall and one
curved wall (see sketch). The shape of the latter may, as a first approximation,
be replaced by two straight-line segments joined a t an angle of 2",as shown in
the sketch.
The Mach Number and pressure are initia!ly 2.0 and 1000 respectively.
Using linearized thcory, make a sketch to scale, showing streamlines, dis-
turbance lines, and pressures in the various regions of constant velocity. Carry
out the flow pattern until two complete reflections of the original wave have
occurred.
14.7. A frictionless, two-dimensional, converging-diverging nozzle is supplied
with air at 100 psia. The diverging section is shaped so that the stream passes
through the exit plane of the nozzle in parallel flow with a Mach Number of 2.0.
The pressure in the exhaust region is variable. Using linearized theory, make a
458 SUPERSONIC FLOW WITH SMALL PERTURBATIONS Ch. 14
sketch to scale of the streamlines and disturbance lines in the region outside the
nozzle exit and indicate the pressures in the various regions of constant velocity,
(a) For an exhaust-region pressure of 14.7 psia
(b) For an exhaust-region pressure of 10.5 psia
14.8. A two-dimensional, supersonic stream of air is bounded by a solid wall
on one side only (see sketch). The wall is shaped so that the stream undergoes
a 20" change in direction along an arc of a circle. It is agreed to replace the
actual wall by a series of straight-line segments of equal length, and it is proposed
to investigate the effect of the number of segments on the calculated values of
the final Mach Number and pressure.
A modified linear theory is to be used, that is, linear theory will be used for
each disturbance line, but the Mach Number downstream of each disturbance
line will be modified so that it corresponds to the newly found pressure.
For an initial Mach Number of 2.0 and an initial pressure of 100 psia, calcu-
late the final Mach Number and pressure when the arc of the circle is replaced
in turn by 2, 4, and 6 equal line segments.
Estimate the asymptotic solution by graphically plotting the results against
the reciprocal of the number of line segments, and compare with the exact
solution (from Prandtl-Meyer theory) of M = 2.83, p = 27.4 psia.
14.9. Considering the wave pattern shown in the sketch, find ( 0 3 - 01)
and Pa - PI as functions of 6, and 6b. Investigate the special cases where
6, = 6 6 and 6, = -61,.
14.10. The sketch shows a two-dimensional duct having an offset formed by
o
l -turns a t the corners A , B, D, and E.
Assuming
-
that a t section AD the stream is uniform and parallel and that
M = 4 2 and p = 100 psia, determine for section CF the way in which the
pressure, Mach Number, and stream direction vary over the cross section, for
values of L/H of 1, 2, and 3.
14.11. The sketch shows a simple form of diffuser for decelerating a super-
sonic stream by very small amounts. Upstream of section AC the flow is uni-
Ch. 14 PROBLEMS 459
form and parallel, and it is desired to design the passage so that the flow down-
stream of section BD will be uniform and parallel.
(a) Show that for small changes in Mach Number, the correct ratio L/H is
related only to M, and find the relationship.
- c and AM.
(b) Find the relation between
(c) Suppose that M = 4 2 , L/H = 0.75, and c = 0.01 radian. Compute
(i) the ratio of the maximum variation in pressure over the section EF to the
average pressure diierence between AC and EF, and (ii) the ratio of the maxi-
mum difference in stream direction over section EF to the wall angle E.
14.12. Consider a supersonic biplane with the lower flat plate always at
zero angle of attack (see sketch). Explore the characteristics of such a biplane
and compare the lift coefficient and center of pressure with the corresponding
parameters for a flat plate. Demonstrate the significance of the parameter
t d ~ , 2 - 1/Z as suggested by Fig. 14.24b.
where p and M are respectively the mean pressure and mean Mach Number in
the test section.
14.16. Show that the lift per unit span for a flat-plate airfoil at small inci-
dence in a supersonic stream is given by
L = - P, VJ'
where 'I is the circulation around the profile.
14.16. The sketch shows a turbine nozzle cut off at an angle of 30" to the
nozzle axis. The nozzle is designed for a Mach Number of g 2 . It operates
with a pressure of 100 psia in the plane shown and discharges into a region where
the pressure is 99 psia.
Ignoring friction, draw a sketch showing the wave pattern and the jet bound-
aries. Indicate compression waves by solid lines and rarefaction waves by
dashed lines. In each zone of flow show the pressure (psia) and flow direction
(degrees). Continue the sketch at least to the point where the jet pattern begins
to repeat itself, and indicate which portion of the pattern repeats itself infinitely.
Air
Moo:
&
E i. .
- .
"
!"
Surface of \
Me
P, = P,
e,, = e,
-
Discontinuity
0
\\ \
\
b
0
and
Chapter 15
(a)
FIG.15.1. Propagation of disturbance produced by bend in wall at point a.
(a) Streamlines and Mach lines according to linear theory. Mean Mach line II,
lies midway between Mach lines of upstream flow (11,)1,and of downstream
flow (I1,)2.
(b) Hodograph diagram according to linear theory.
(c) Exact Prandtl-Meyer solution for corner flow.
that these lines are also lines of constant perturbation potential, and
that the vector change in velocity produced by a pressure wave must
accordingly have a direction normal to the Mach direction. From these
considerations, all the calculations can be reduced t o a single unit proc-
ess, illustrated by Fig. 15.1.
In Fig. 15.1a1an initially uniform parallel flow in region 1 is influenced
by a small turn in the wall a t point a. The effect of this turn is propa-
gated along the left-running Mach line II,, which makes the Mach
angle al to the initial flow. Fig. 15.lb shows how the velocity V 2 is
determined from the known turning angle 6 and from the rule that the
vector change in velocity is normal to the line II,. The inclination of
the latter depends on the Mach Number. As a first approximation
the Mach line I I , may be drawn a t the Mach direction corresponding
to the upstream flow, i.e., as line However, after V 2 and other
stream properties are tentatively found in region 2 by means of this
first approximation, the accuracy of the approximation may be improved
by taking the line IZ, a t the mean Mach angle E t o the mean flow direc-
tion. In this way, by several iterations, a better value of the stream
properties in region 2 may be computed with the help of Fig. 15.lb, a n d ,
the direction of the mean Mach line I I , may be more accurately drawn
in the physical plane. Fig. 15.la.
These ideas may be extended to the flow around the curved wall of
Fig. 15.2a. As an approximation, the continuous. wall is replaced by a
"a
/
I- /
/
3' .1
.-
(a) (b)
FIG.15.2. Flow past curved wall by modified linear theory.
(a) Streamlines and Mach waves.
(b) Hodograph diagram.
-- \, -du
V t dV
P
(a) (b)
FIG.15.3. Infinitesimal Mach wave.
(a) Physical plane.
(b) Hodograph plane.
Now, from the adiabatic relations of Chapter 4, khe relation between dT'
and d~ is
6 =-
k-1
+ constant
arc tan ,/r (M' - I ) + arc tan 7
M' - 1
(15.4)
It is often more convenient to work with the dimensionless velocity
M* = V / c * rather than the Mach Number M. Using the adiabatic r e
lation that
RIGHT-RUNNING
WAVE. If we had been dealing with a problem in
which there were Mach waves only of family I, the sign on the right-
hand side of Eq. 15.1 would have been reversed, and we would have
ended with
BIT = +w(M*) f 211 - 1000 (15.7)
in the hodograph plane; and, with simple right-running waves, the Mach
line a t a given point is normal to the hodograph characteristic of family
I I a t the image point.
the right-running Mach lines lie a t the angle (0 - or), measured counter-
clockwise.
Many different simple-wave flow patterns may have the same image
in the hodograph diagram. For example, Fig. 15.8 shows the flo-NSalong
three curved walls, all with trhe same initial Mach Number, and all of
which are represented by the same hodograph characteristic. At each
point of each wall the flow properties are determined uniquely by the
streamline direction, and these flow properties are propagated out from
the wall along the corresponding Mach lines.
472 METHOD OF CHARACTERISTICS Ch. 15
\
f
\
Mg
(0)
(dl
Fra. 15.8. Three simple-wave flows represented by the same hodograph charaoter-
istic.
(a, b, C) Physical plane.
(dl Hodograph plane.
473
474 METHOD OF CHARACTERISTICS Ch. 15
(4 (4
FIG.15.9. Centered simple waves.
15.8 that the maximum possible turning angle of the flow is also 130.5".
With this turning angle the stream would expand to zero pressure and
temperature. Since orl = SO0 and a2 = 0, the tangent to the hodograph
characteristic is along the radius a t the M* = 1 circle, and is normal to
the radius a t the M*,,, circle (Fig. 15.10~). This may also be deduced
from Eq. 15.3.
Query: what is the flow pattern when a flow with Mach Number unity
approaches's corner which turns through an angle of more than 130.5",
(4
FIG.15.10. Complete simple wave.
(a)
I
I
,
I
11'
rI \
/ '
. ,
(b)
0 40 80 120 160
v.90- ( a - w ) , Degrees
(c) (dl
FIG.15.11. Construction of streamlines for simple wave.
(a) Noncentered simple wave.
(b) Centered simple wave.
(c) Hodograph diagram.
(dl Chart for facilitating construction of streamlines.
sectional area for isentropic flow, and the distance h sin a is proportionai
to the cross-sectional area for flow a t any other section. We may there-
fore write
(h sin a)/h* = A/A*
where A/A* is the area ratio for isentropic one-dimensional flow, related
to the Mach Number by the formula (Chapter 4)
and the streamlines are easily plotted with r and (a- w ) as polar
coordinates.
Thus, given 6' and w , 1 and 11 may be found from Eqs. 15.12. Or, con-
versely, given I and 11, 0 and w may be found from Eqs. 15.13. The
latter relations show also that in the hodograph plane, radial lines
(lines of constant flow direction) through the origin are lines of constant
+
(I 11), whereas concentric circles (circles of constant velocity and
pressure) are lines of constant ( I - 11).
Art. 15.3 EXTENSION OF LINEAR THEORY 479
Hence, from Table B.7,w3 = 20". Then, since states 1 and 3 lie on a common
I-characteristic, we have 1 3 = 11= 505. Eqs. 15.12 and 15.13 then give
03 = -10" and 1 1 3 = 485. Thus, there is a 10" deflection of the jet boundary
as it leaves the nozzle.
All the details of the corner flow between fields 1 and 3 are known analytically
from the simple-wave formulas of the preceding article. Subsequently, when
waves of both families are present, however, only stepwise methods can be used.
Therefore, we replace the actual Prandtl-Meyer continuous expansion by a
number of individual Mach waves into which the actual changes of state are
lumped. For the sake of illustration, we have here substituted two individual
waves, a and b, for the actual corner flow. Assuming equal angles of turn across
each of these waves, we have O2 = &/2 = -5O and 1 2 = I1 = 505. Then
w2 = 15.0 and ZZz = 490.
Points l', 2', and 3' may be spotted in the hodograph diagram of Fig. 15.12b.
Wave a is then drawn in the direction a', i.e., normal to the hodograph char-
acteristic of family I at the mean conditions between fields 1 and 2. Similarly,
wave b is drawn in the direction b', normal to the hodograph characteristic of
family I at the mean conditions between fields 2 and 3.
When wave a strikes the upper wall, it must be reflected as waye c in order to
have the flow follow the upper wall. If the region downstream of wave c is
called field 4, we have O4 = 0 and 1 1 4 = 1 1 2 = 490. Hence 1 4 = 510 and
wa = 20.0. Wave c is now drawn in the direction c', normal to the hodograph
characteristic of family I1 at the mean conditions between fields 2 and 4. The
intersection of wave c with wave b defines one corner of the Mach quadrilateral
for field 2.
Now assume that wave b is continued as wave d, and that wave c is continued
as wave e, and that they partially bound field 5. But field 5 is reached from
field 4 along a characteristic of family I, and is reached from field 3 along a
characteristic of family 11. Therefore Z6 = Zp = 510, and 1 1 6 = Z13 = 485,
whence ws = 25.0 and 196 = -5.0. Waves e and d are now drawn with the
respective directions e' and d'.
What happens when wave e strikes the free boundary of the jet? Since field 5
has a lower pressure than the surroundings, wave e must be reflected in the form
of wave g'to preserve the boundary condition of constant pressure. If the field
480 METHOD O F CHARACTERISTICS Ch. 15
downstream of wave g is 7, we have 0 7 = 0 3 = 20.0 and 17 = 1 5 = 510, whence
117 = 490 and 0, = 0. Wave g may now be drawn with the direction g', which,
incidentally, coincides with the direction d'. The lower side of field 7 is the jet
boundary, which is drawn with the appropriate streamline direction, 07 = 0.
remaining two properties in each field are of course found from the two under-
lined ones either algebraically from Eqs. 15.12 and 15.13, or graphically from
Fig. 15.12b.
Field I I1 o e a e + a e P/PO
+
The table also shows in each field the direction (0 a) for left-running waves,
and (0 - a) for right-running waves. These may be combined as in the "Table
for Fig. 15.12a" to get the inclinations of each wave. With the help of this
table, Mach wave directions may be determined without graphical operations
in the hodograph plane.
482 METHOD OF CHARACTERISTICS Ch. 15
characteristics are identical with the Mach lines of the flow. The left-
running Mach lines are physical characteristics of family I1 and the
right-running Mach lines are physical characteristics of family I.
These results may be seen even more easily by arbitrarily placing the
z-axis parallel to the velocity vector a t a given point in t h e flow field.
At that point we would have u = V, v = 0, M = V/c, and 0 = 0 ;
Eq. 15.20 then becomes
lTIZ
Since, a t the point in question, the velocity vector coincides with the
x-axis, we conclude that the physical I-characteristic lies a t the Mach
angle below the velocity vector and the physical II-characteristic lies at
the Mach angle above the velocity vector.
Orthogonality Relations Between Physical and Hodograph Character-
istics. I t may be verified by algebraic simplification, using Eqs. 15.20
and 15.21, that
which means that the physical characteristic of one family is normal to the
hodograph characteristic of opposite family. This rule of reciprocal ortho-
gonality between the hodograph and physical characteristics is illus-
trated schematically in Fig. 15.13, and may be used to advantage for
graphical solutions.
Equations of the Hodograph Characteristics. The characteristic
curves for any particular problem are found by simultaneous integration
of Eqs. 15.20 and 15.21. Once these curves are established, the entire
problem is solved, for the characteristic curves relate points in the
physical plane to points in the hodograph plane, and hence the-Mach
Number and flow direction may be found for each point of the physical
flow field.
Now Eq. 15.20 shows that the differential equation of the physical
characteristics contains velocity terms, and must therefore be solved
simultaneously with the differential equation of the hodograph charac-
teristics. Quite differently, however, the differential equation (Eq.
15.21) of the hodograph characteristics contains only velocity terms, and
may therefore be integrated in advance of any specific physical problem
to give a set of universal hodograph characteristics which are independ-
ent of the flow pattern in the physical plane.
486 METHOD OF CHARACTERISTICS Ch. 15
In obtaining the universal set of hodograph characteristics, it is con-
venient to employ polar coordinates, V, 8, rather than Cartesian coordi-
nates u, v. Applying the transformation of Eq. 15.22 t o Eq. 15.21, we
get after algebraic manipulation,
1
= 7 tan CY (15.25)
An even simpler method of getting Eq. 15.25 is to use the rule which
states that the normal to the hodograph 1-characteristic lies a t the angle
+a tu the velocity vector, while that for the 11-characteristic lies a t the
1 a . Applying this rule to the graphical construction of Fig. 15.14,
Eq. 15.25 becomes immediately
1
a
Eq. 15.25 is seen to be identical
th Eq. 15.1. Therefore, making
the substitution of Eq. 15.2, and
carrying out the integration of
Eq. 15.3, we arrive a t
01 = -w(M*) + 2 1 - 1000
(15.26a)
u 811 = +w(M*) + 211 - 1000
FIG.15.14. Geometrical interpretation
of the differential equation of the hodo- (15.26b)
graph characteristics.
where w(M*) is given by Eq. 15.6
and Fig. 15.4. Obviously Eqs. 15.26, representing the hodograph charac-
teristic curves, are identical with Eqs. 15.5 and 15.7. The two families
of curves of Fig. 15.5 are in fact the unique set of hodograph charac-
teristics which apply to any problem of two-dimensional, supersonic,
irrotational, isentropic flow. The symbols 1 and 11are the character-
istic coordinates, related to the state coordinates, w and 8, through Eqs.
15.12 and 15.13.
Example of Construction of Characteristics Nets. I t has already been pointed
out that, as a consequence of the special properties of the characteristic curves,
the original problem of finding a solution ~ ( xy), of Eq. 15.14 for given initial
conditions is replaced by the equivalent problem of constructing in the physical
and hodograph planes the characteristic curves of Eqs. 15.20 and 15.21. In a
sense this procedure gives the desired kinds of answers more directly, because
it is the velocity vector rather than the velocity potential which is of interest.
,The hodograph characteristics for all problems are given by Eq. 15.26. To
solve problems having specific physical boundary conditions, we construct
stepwise in the physical plane a characteristics net of whatever degree of fineness
is necessary to obtain the desired accuracy.
Art. 15.4 APPLICATION O F T H E O R Y O F CHARACTERISTICS 487
Lattice
Point I I1 w 0 a 0+a 0-0 P/PO
Since the hodograph characteristics are known to begin with, points 5', 6',
and 7' may be located immediately in the hodograph plane by means of the
hodograph characteristics passing through l', 2', 3', and 4'. Or, since the values
of I and II for 5', 6', and 7' are determined by the values of 1and 12for l', 2',
3', and 4' (e.g., 1 6 = I I and I I g = IIz),Eqs. 15.12 and 15.13 may be used for
filling in the values of w and 0 in the attached table.
To find point 5 in the physical plane, the physical I-characteristic 1-5 is drawn
normal to the hodograph 11-characteristic at the mean conditions between 1'
and 5'. Similarly the physical 21-characteristic 2-5 is drawn normal to the
hodograph I-characteristic at the mean conditions between 2' and 5'. The inter-
section of the two physical characteristics establishes the location of point 5.
If it is not desired to use the hodograph curves, wave 1-5 may be laid off at the
appropriate mean inclination (0 - a),and wave 2-5 a t the appropriate mean
+
inclination (0 a);these inclinations are shown in the attached table.
In locating points 8 and 8', we note that point 8 is on the physical boundary
and that the velocity direction a t 8' is known. This, together with the fact
that 8 lies on the 11-characteristic passing through 5, establishes points 8 and 8'.
Points 11 and 11' are found in the same way.
By continuing this process, the characteristics nets are established in both
planes. In the physical plane of Fig. 15.15a they are shown as solid lines con-
nected by small circles at the net points.
Figs. 15.15~and 15.15d show that the Mach wave pattern as determined by
this method is in good agreement with the pattern observed in schlieren photo-
graphs.
and
490 METHOD O F CHARACTERISTICS Ch. 15
where either the positive or negative sign, but not both, is applicable,
because u and v are uniquely related. By comparison with Eq. 15.21,
it is seen immediately that the hodograph curve representing the unique
relation between u and v must in fact be one of the hodograph charac-
teristic curves.
Next, using the equation of irrotationality and the fact that u = u(v)
we carry out the following operations:
so that dv
-- - -l/(d~/dx)~
du
Now dv/du is, as we have just seen, the slope of the hodograph charac-
teristic. Also, ( d y l d x ) , is the slope in the physical plane of lines of con-
stant v, and therefore of constant fluid properties, since u depends only
on v. Recalling now the reciprocal orthogonality relationships of Eq.
15.24, the equation immediately above proves that in a simple-wave
flow the lines of constant fluid properties are Mach lines of opposite
family to the hodograph characteristic on which the simple-wave flow
is mapped. Furthermore, since u and v are constant on each such Mach
line, the stream direction and Mach angle are likewise constant on each
Mach line. Hence the Mach lines under consideration must be straight
lines.
Art. 15.6 FIELD METHOD VERSUS LATTICEPOINT METHOD 49 1
(a) (b)
FIG.15.16. Cancellation of a wave by appropriate curvature of wall.
and thus x3 and y3 may be solved for in terms of the initial data and the
values of bI and bII.
ORGANIZATION OF NUMERICAL CALCULATIONS. There are several ways
of organizing the calculations so as to put them on a routine basis suit-
able for computers. One type of organization, t o be used with the lat-
tice-point method, is illustrated in Fig. 15.18. The Mach net is twice
(sj, 0-4
FIG.15.18. Scheme of organization for purely numerical solution.
(a) Schematic characteristics net; computation of fluid properties.
(b) Schematic chvracteristics net; computation of coordinates of net points.
At the boundary of the physical flow, one of the lines in Fig. 15.17
will be a segment of the boundary, and a modification of the computa-
tional technique is required. Either the flow direction or the fluid pres-
sure is then known along one of the lines in Fig. 15.17, and equations
similar to Eqs. 15.27 and 15.28 may be worked out for these special cases.
Effect of Grid Size on Accuracy. There is only one certain way of
determining the required Mach net for a specified final accuracy, and
that is to start with a very coarse net and obtain solutions with nets
successively finer, until two successive solutions agree t o the desired
accuracy. Occasionally it is possible to extrapolate to zero grid size.
A convenient rule is that the error owing t o the finite-difference type of
computation is proportional t o ( ~ w ) where
~, Aw is a measure of the
change in M* between fields or between lattice points.(3)
For example, Fig. 15.19 shows, for the sharp-cornered supersonic wind
tunnel nozzle of Fig. 15.34, the computed values of the exit coordinates
(4 (b)
FIG.15.22. Flow past thin profile.
(a) Flow plane.
(b) Hadograph plane.
joined to each other, however, only along Mach lines, which are the
patching lines for supersonic flow. On the Mach lines the velocity is
continuous, but there are discontinuities in the first and higher deriva-
tives of the velocity.
The pattern shown in Fig. 15.23 is repetitive, since field 5 is identical
in state with field 1.
j -i7--
82-g3,
I f 500 1
(b)
Fro. 15.24. Design of a simple supersonic wind tunnel.
(a) Flow plane.
(b) Hodograph plane.
__------
the interaction between the resulting shock wave and the boundary
1a;yer makes the situation grow rapidly worse. There are likely to be
great disappointments, therefore, in the results obtained with passages
like those of Figs. 15.25, 15.26, and 15.27 unless due account is taken of
these problems. Shocks will usually appear and the general flow pattern
will be totally unlike what is expected. There is evidence, however,
that good results may be obtained with decelerating flows if three pre-
cautions are taken: (i) very large adverse pressure gradients should be
avoided, (ii) the passage walls should be displaced by the amount of
a carefully computed boundary-layer displacement thickness, and (iii)
any large initial boundary layer should be removed by suction. For
further details, the reader should refer to Article 16.12 and to Volume
11, Chapter 28.
Test S e c t ~ o n
[Uniform, Parallel
Flow 1
(a)
/
A\----
\
..'4
boundary of the flow for calcula-
W~.
/
:1---.
.,L -.
/- . -7
.-
tion purposes. Hence we need
consider only the upper half of
the nozzle.
e, T e = 0 5 ~ 7 * 4' I,,?' In order to have a net increase
e,=~,
in area without any net change in
(b) flow direction, the wall contour
FIG. 15.29. Nozzle for producing m i - must first curve outwards from
form, parallel flow.
7 to 3, and must then curve in
(a) Flow plane.
(b) Hodograph plane. again until a t the exit (point 1)
the wall is parallel to the initial
flow. Point 3 is an inflection point and is the point where the wall has
its maximum slope, Om,,. For the present we shall assume that the
subsonic inlet is designed to give a parallel, uniform, sonic flow a t the
throat, section 6-7.
ZONES I N NOZZLE. Three separate zones of flow may be identified in
Fig. 15.29a:
(i) The expansion zone, 6-7-3-2-6, bounded by the throat 6-7, by
the characteristic 2-3 which on the center line of symmetry attains the
required test-section Mach Number, and by the expansion portion of the
nozzle wall 7-5-3. This is a zone where waves of both families are
present.
(ii) The test section, which is the region downstream of 1-2, where the
flow is uniform and parallel a t the test-section Mach Number ME. Be-
Art. 15.11 SUPERSONIC WIND TUNNEL NOZZLES 509
cause of the uniform, parallel flow, the Mach line 1-2 is straight, and is
inclined a t the angle c u ~to the center line.
(iii) The straightening section, 3-2-1, which is bounded by the Mach
lines 3-2 and 2-1, and by the straightening portion of the wall, 3-1.
Since it is a general theorem that only a zone of simple waves can be
patched to a uniform, parallel flow, it follows that zone 3-2-1 must
have a simple-wave flow such that all flow properties are uniform on
straight left-running Mach waves in this zone. The wall contour is then
the streamline which begins a t point 3 and crosses the left-running
Mach waves until it emerges horizontal a t point 1. Since it is undesirable
for reasons of boundary-layer behavior to have compression waves in
the nozzle, the simple wave in the straightening section must be an ex-
pansion wave, and hence the straightening contour 3-1 must be concave
downward. Note that the determination of the flow in this zone from
initial data on the Mach lines 3-2 and 2-1 is an initial-value problem
of the type of Fig. 15.20b.
We may also discuss the main principles of the design using the physi-
cal concept of pressure waves propagating along the Mach lines. Assum-
ing sonic flow a t the throat, the wall is curved outward between 7 to 3.
This generates right-running expansion waves which are reflected from
the center line of symmetry as left-running expansion waves, which may
be reflected again and again a t the wall and the center line.
This process of generating expansion waves by curving the wall is
continued to the point where one of the waves (3-2) accelerates the flow
a t the center line (point 2) to the desired test-section Mach Number.
The left-running waves which cross line 3-2 would in general be reflected
from the wall. However, the purpose of the design is to avoid pressure
waves in the test section, and hence the straightening contour 3-1 is
curved so as to cancel the waves which are incident on it. Thus the re-
gion 3-2-1 is one of simple waves.
It is clear that if the expansion contour 7-5-3 is selected arbitrarily,
thus allowing the Mach line 3-2 to be constructed, it is possible then to
construct the streamline 3-1 and thus complete the nozzle design. We
now ask, what are the considerations that go into the selection of the
expansion contour, 7-5-3?
Design of Expansion Contour, 7-5-3. The design of the expansion
contour depends somewhat .on the shape of the sonic line a t the throat.
For illustrative purposes, and to make the discussion concrete, let us
suppose that the sonic line is straight and normal t o the axis, as in
Fig. 15.29a.
Furthermore, also for the sake of concreteness, let us assume for the
present that the first test-section Mach line (2-I), when extended back-
ward, is reflected a t the wall contour twice (at points 3 and 5) before
510 METHOD OF CHARACTERISTICS Ch. 15
reaching the sonic line. Now, if the radius of curvature of the contour
7-5 were finite, the waves originating a t 7 would be reflected infinitely
many times a t the axis and the contour before the Mach Number in-
creased above unity, because the Mach line running rightward from 7 is
normal to the axis, and likewise the Mach wave reflected from 6 is also
normal to the axis. Therefore, to meet the stipulation that the back-
wardly extended Mach line 1-2 be reflected only a t 3 and 5, it is neces-
sary that the contour 7-5 have zero radius of curvature. This means
that points 5 and 7 coincide in physical location, that there is a corner
a t 7, and that the lines 6-7 and 5-7 are identical.
Although in the triangular region 7 - 4 4 there are waves of both
families, the corner 7 is a singular point a t which there is locally a cen-
tered simple expansion wave, so that the series of states between 7 and
5 map on a single hodograph characteristic, as in Fig. 15.2913. The lat-
ter figure also shows the hodograph characteristics corresponding to the
backwardly extended Mach line 1-2.
Now, from Eqs. 15.26, since B6 = Bp = el = 0,
Furthermore, since U6 = 0,
Therefore,
wl = 193) + le,l + b5l 4- lO5l = 2l9,l + 21e51
1931
- - bsl -
- 1 - ---- W5
-1-2- (15.29)
4 2 4 2 0 1
Thus, equations 15.30 and 15.31 put certain limitations on the design.
It is instructive to examine the two extremes:
Art. 15.11 SUPERSONIC W I N D TUNNEL NOZZLES 511
(i) 4 0 ,= 1/4 and 1831 = ul,!4. In this case, = 10.71,SO that the
contour 5-3 is a straight line whose slope is related to the test-section
Mach Number by
1031 = lesl = 01/4
The corresponding nozzle contour and characteristics diagram are shown
in Figs. 15.30a and b.
-
Test
-.
6 4
('J)
FIG.15.30. Nozzle with w b / w l = % and 1031 =4 4 .
(a) Flow plane.
(b) Hodograph plane.
(ii) 05/01 = 0 and 14= w1/2. In this case, points 5 and 7 are iden-
tical as to both physical location and fluid properties, but, since a cor-
ner a t 7 is still necessary, point 3 is coincident with 7, although it has
different properties. The maximum slope of the contour is immediately
(b)
FIG. 15.32. Nomenclature for generalized nozzle-design relations.
corner, let B denote the infle tion point of the contour, and let EE r e p
resent the first Mach line in the test section. Then
The construction of the wave pattern and wall contour for such a
nozzle, using the field method, is illustrated graphically in Figs. 15.33a
and b.
7'
A scale drawing of the Mach net for W E = 28' (ME = 2.059, BB = 14O),
using lo-intervals, is shown in Fig. 15.34. I t is to this design that Fig.
15.19b, illustrating the effect of Mach-net fineness on accuracy, pertains.
No nozzle not conta.ining compression waves can have a smaller ratio
of length to throat width nor a larger value of 8,- than the sharp-corner
FIG. 15.34. Mach lines for sharp-corner nozzle with exit Mach Number of 2.059
(WE = 2g0), based on lo-intervals of w . The origin of the x, y-coordinate system is
in the plane of the throat, on the line of symmetry, and y~ denotes the half-height
a t the throat (M.I.T. Gas Turbine Laboratory).
The region with waves of both families (zone 74-6 in Fig. 15.31), if
drawn for a very high exit Mach Number, contains within it the mixed-
wave zone for all lower Mach Numbers. This is illustrated in Fig. 15.35,
Contour For w~ = I 2
------r---
-.
FIG. 15.35. Illustrating construction of contour for sharpcorner nozzle from "kernel."
0.6
0 Measured
Theoretical
FIG. 15.37. Length of sharp-corner noz- FIG. 15.38. Comparison between meas-
zle (after Shames and Seashore). wed and theoretical pressure distribu-
tion on axis of sharp-corner nozzle of
Fig. 15.34 (after Edelman).
Since we are considering regions of flow from which shocks are ex-
cluded, the entropy is a continuous function of x and y. Hence
and the partial derivatives ap/ax and a p / d y which appear in Eqs. 15.36
and 15.37 may therefore be written
a ~ / a z= (PC +
/ k ) ~ ( s / c , ) / ~ x c2 a p / a x (15.41)
aday = ( P C ~ / ~ ) ~ ( s / c , ) / ~+
Y c2 splay (15.42)
Since u, v, p, and s are all continuous functions of x and y, we may
write
+
du = ( a u / a x ) d x (aulay)dy (15.43)
dv = (av/ax) d x + (av/ay) d y (15.44)
dp = (ap/ax) dx + ( ~ P / ~ d~Y I (15.45)
= (as/ax) dx + (aslay) d y (15.46)
Solution of Equations. The expressions for dp/dx and a p / a y given by
Eqs. 15.41 and 15.42 are now substituted into Eqs. 15.36 and 15.37,
respectively. We then have eight quantities, namely, au/ax, du/ay,
dv/ax, av/ay, ap/ax, ap/ay, &/ax, and aslay, which are connected by
eight independent, linear, nonhomogeneous equations, namely, Eqs.
15.35, 15.36, 15.37, 15.38, 15.43, 15.44, 15.45, and 15.46. The solution
for each of the eight quantities may be written symbolically in terms of
the augmented determinant of the eight equations:
it is clear that Eq. 15.4% represents Euler's equation for changes in-
state along a streamline. Hence, Eqs. 15.48 state in sum that the
streamlines are characteristic curves in a rotational adiabatic flow.
Physically, this means that the streamlines are possible slip lines, or
vortes sheets in an infinitesimal sense, on which there may be disconti-
nuities in the derivatives of the velocity, density, entropy, and pressure.
If Eq. 15.49 is solved for dy/dx, the resulis will be recognized as being
identical with those of Eq. 15.20. Hence Eq. 15.49 gives the differential
equations of the physic31 characteristics, and can be reduced to the form
of Eq. 15.23, namely,
This shows that these physical characteristics make the Mach angle t o
the velocity vector and are, therefore, identical with the Mach lines of
the flow.
Turning to Eq. 15.50, it is convenient to express the velocity vector in
terms of polar coordinates V and 8, by means of the relations u = V Cot3 e
and v = V sin 8. Making this transformation, and eliminating dy/&
520 METHOD OF CHARACTERISTICS Ch. 15
from Eq. 15.50 with the help of Eq. 15.51, we get the differential equa-
tion of the hodograph characteristics in the form
dB dp
3=
sin a! cos a
- +- + 1- ~ ( s / c , )= 0
p k
dp - 1 ds VdV
- - - -. ---
P k-lc, c2
Substitution of Eq. 15.53 into Eq. 15.52 yields, after rearrangement,
7df.I----
1 dV sin a cos a!
d
(;-") =O (15.54)
tan a! V k(k - 1 )
This may be solved for dV/dO to give
By comparing Eq. 15.55 with Eq. 15.25 we see that the effects of
vorticity are contained in the third term of Eq. 15.55. This term depends
in part on the physical variables x and y, because the entropy is deter-
mined by the streamline rather than by the Mach Number. Conse-
quently it is not possible to integrate immediately the hodograph charac-
teristics, as was possible for irrotational flow. Instead, for each specific
problem, both the physical and hodograph characteristics must be
simultaneously constructed in a stepwise manner. Another consequence
of the presence of the third term in Eq. 15.55 is that i t eliminates the
possibility of there being regions of simple waves in a rotational flow.
Numerical Method of Solution. The solution of practical problems is
carried out by constructing the physical and hodograph characteristics
nets piecewise by simultaneously solving Eqs. 15.51 and 15.55 in finite-
Art. 15.12 FLOW W I T H ROTATION 521
difference form, replacing an element of arc on one of the characteristic
curves by a straight-line chord. Several related methods for carrying
out such calculations are described
in detail in Reference 3. A brief ,
outline of one line of attack is
given below in terms of the unit
process required to construct the
characteristics net.
Suppose that a t points 1 and 2
in Fig. 15.40 we know X I , yl, 22,
Y 2 , v 1 , v 2 , 81, 8 2 , S l , s 2 , thus corn-
pletely defining the location and
fluid properties of these two
points. Let point 3 lie at the in-
tersection of the physical ZI-char-
acteristic passing through point 2
and of the physical Z-characteris-
tic passing through 1. The prob-
2 -X
FIG.15.40.Determination of properties
lem is to determine the bcation a t point 3 from data a t point*^ 1 and 2.
and the fluid -ropert - ties of point 3.
Equations 15.51 may be written in finite-difference form
and V are uniquely connected with each other. Moreover, the entropy
depends only on the particular streamline on which a given point lies.
Eqs. 15.56 to 15.59 may be solved simultaneously by an iteration
method. The values of V3 and 8 3 are first chosen tentatively. Using
these values, simultaneous solution of Eqs. 15.56 and 15.57 yields ap-
522 METHOD OF CHARACTERISTICS Ch. 15
proximate values for x3 and ys By extrapolating the streamlines toward
point 3, the entropy of point 3 may be tentatively determined. Then
Eqs. 15.58 and 15.59 may be solved simultaneously for V3 and 83. If
these values depart too greatly from the values originally assumed, the
process of calculation is repeated again and again until satisfactory
convergence is obtained.
Successive application of this type of calculation allows the entire
characteristics net to be determined in a manner exactly like that for
irrotational flow, with similar considerations applying as to the way in
which various types of initial data determine the regions in which solu-
tions may be obtained.
,
Semigraphical Method of Solution. The tediousness of the iteration
method of calculation can be partly relieved by employing a semi-
graphical procedure which employs the irrotational hodograph charac-
teristics. Noting that
dV/V = dM*/M*; sin a = 1/M; tan a = I/-
and using the adiabatic relation between M and M*, we may put Eq. 15.55
in the form
(15.60)
Expressed in finite-difference form, this becomes
i
k-I- k-l-
I--- M * ~ I--- M *2
(AM*)I, 11 = TM*
Ic +1 A0 -
k+l
2k(k - I ) -
M *
- -
~ 1 M*
k+l
(15.61a)
where M* refers to the mean va!ue of M* over the interval to which
Eq. 15.61a is applied. This equation is also expressible as
Returning now to the unit process of Fig. 15.40, we may from the
given initial data locate the image points 1 and 2 in the hodograph dia-
gram of Fig. 15.41. The mean Mach lines I and II are now tentatively
drawn in Fig. 15.40, thus establishing an approximate position for
point 3 and allowing the entropy a t 3 to be determined tentatively.
The second term on the right-hand side of Eq. 15.61b may then be
approximately calculated for each characteristic, and the correspond-
ing partial values of AM* laid off radially in Fig. 15.41 as the lines 1-la
Compare the exact theory with the linear theory by plotting curves of p,/p,
versus MI for two values of 6, namely, 2O and 10". Use a range of M1from 1 to 3:
15.2. Compare supersonic flow past a convex corner with (a) incompressible
flow, (b) subsonic flow. If the fluid is compressible, is it possible to have a purely
subsonic flow in the neighborhood of the corner? How would a boundary layer
near the wall influence the supersonic flow past a convex corner?
15.3. Plot a Prandtl-Meyer streamline from M1 = 1 to Mt = 5.
15.4. I n simple-wave flows what are the maximum stream turning angles for
(a) Decelerating flows with initial Mach Numbers of (i) 1.0, (ii) 2.0, (iii)
5.0, and (iv) 10.0?
(b) Accelerating flows with initial Mach Numbers of (i)' 1.0, (ii) 2.0, (iii)
5.0, and (iv) 10.0?
Ch. 15 PROBLEMS 525
15.5. A two-dimensional jet leaves a convergingdiverging nozzle in parallel
flow (see sketch) and discharges into the atmosphere, where the pressure is
14.7 psia. The area ratio of the nozzle
is 2.0.
Calculate the angle 6 in degrees, if
(a) po = 200 psia
(b) po is infinite
15.6. Suppose that in Fig. 15.118
the shape of the wall is given by
Y = f(x1, Yl, 4
Find the equation of the streamline
passing through point l', in terms of
XI,, yl,, and the functional relation PROB. 15.5.
above.
Generalize your results to any arbitrary initial Mach Number, introducing
MI as a parameter.
15.7. Consider a corner-type flow beginning a t Mach Number unity. .Derive
the following equation in polar coordinates for a streamline:
where r is the radius vector to the streamline; r* is the radius vector to the same
streamline a t M = 1; and v is the angle measured clockwise from r* to r .
15.8. Using the theory of characteristics as outlined in Appendix A, deter-
mine the shapes of the characteristic curves in the physical and hodograph
planes for the differential equation corresponding to supersonic, uniform,
parallel flow with small perturbations.
15.9. Consider a perfect gas with k = 1.0
(a) Investigate the shape of the hodograph characteristics.
(b) For a complete corner-type flow determine in polar coordinates the equa-
tion of a streamline in terms of r , the radius vector to the streamline; r*, the
radius vector to the same streamline a t M = 1 ; and v,-the angle measured clock-
wise from r* to T .
15.10. Show that a flow region adjacent to a region of uniform, parallel flow
must be a region of simple waves.
15.11. (a) Demonstrate that when a pressure wave is reflected from a plane
wall,
(b) Show that when two pressure waves of opposite family penetrate each
other, the values of A0 and of Aw for each wave are the same after penetration
as they are before penetration.
(c) Derive a corresponding rule for reflection of a wave from a constan[-
pressure boundary.
526 METHOD OF CHARACTERISTICS Ch. 15
15.12. (a) Suppose that, in Fig. 15.17, the line 2-3 is a solid boundary, and
that 0 is known on this boundary. Work out the computation formulas for
determining conditions a t 3 from those a t 1 and 2, given wl, 01,xi, yi, oz,02,
xz,y2, and given also 83 and ya as functions of x3. Note that iteration is required.
(b) Suppose that in Fig. 15.17, the line 2-3 is a constant-pressure boundary
of the flow, as for a free jet. Work out the computation formulas for determining
conditions a t 3 from those a t 1 and 2, given wl, 01,XI, yl, WZ,02,52,y2, and wa.
15.13. Demonstrate that if in a given region of the flow all states are mapped
on a single hodograph characteristic of family I , then in that region all the
left-running Mach lines are straight lines, and on each left-running Mach h e
all fluid properties are constant.
Are the right-running Mach lines straight lines for this case?
15.14. Consider a plane in which the Cartesian coordinates are the charac-
teristic coordinates, Z and ZZ (see sketch).
(a) Sketch in this plane lines of constant Mach Number and of constant flow
direction.
(b) Plot the fields of Fig. 15.128 in this plane.
(c) Plot the net points of Fig. 15.15a in this plane.
15.15. A thin profile has the shape of a parabola, with its leading edge parallel
to the oncoming air stream (see sketch).
Assuming that t / l = 0.10 and M l = 2.059, find the lift and pressure drag
coefficients.
<
15.16. A nozzle is to be designed to deliver a parallel, uniform stream of a,ir
a t a Mach Number of 2.059. The general arrangement of the nozzle is as shown
in the sketch.
- FI,
,;' it.'
The inlet section will be assumed to have a shape which produces a hori-
zontal parallel flow a t the throat section AB.
Ch. 15 PROBLEMS 527
In the region AFDB the flow is an expansive "corner flow" around the
corner C. The length AC is twice the length AB.
In the region FED the flow is uniform and parallel.
In the region FHIE the flow is an expansive "corner flow" around the corner
G. The length GE is twice the length FE.
Region J is a uniform, parallel flow with a Mach Number of 2.059. The
direction of the stream at J is the same as the direction at the throat AB.
Assuming that the throat height AB is one inch, find the lengths and inclina-
tions to the horizontal (in degrees) of the following lines: FD, FE, DE, and HI.
Using these principal dimensions and directions, make a sketch to scale of the
nozzle.
16.17. Work out the problem of Fig. 15.12 by the lattice-point method.
16.18. Consider a two-dimensional, source-type flow with an initial Mach
Number of 1.0 and a final Mach Number of 2.0. Using the method of char-
acteristics, plot pressure versus radius for stepwise solutions with turning angles
across each wave of 8°, 4°, and 2°, and compare with the exact solution, corre-
sponding to a turning angle across each wave of 0°.
16.19. Design a symmetrical blade for an impulse turbine with a turning
angle of 120° and an entering and leaving Mach Number of 2.059.
16.20. Design a stationary supersonic compressor cascade of the type shown
in Fig. 15.27, assuming that M1 = 2.059, P1 = 10, that the air entering is in-
clined 60° to the axial direction, that a normal shock occurs at M = 1.4, and that
the flow leaves in the axial direction.
Lay out a pair of blades, and determine the leaving Mach Number, pressure,
and the over-all ratio of stagnation pressures.
16.21. A uniform, parallel flow with M1 = 2.059 and P1 = 100 approaches
the two-dimensional 90°-bend shown in the sketch. Determine the flow pattern
leaving the bend.
~~~--- 2h - - - - - ;
1
The design is to be based on sharp corners a t the throat section aa', a t which
section it may be assumed that the flow is uniform and parallel. The contours
a'b'c'd' are to be designed so as to produce a uniform, parallel stream in region 4.
In carrying out the design, use the wave pattern suggested in the sketch, selecting
the Mach lines between regions 1 and 2 to correspond to equal deflection angles
across each of the Mach lines.
Assuming that the height of the nozzle a t the throat (distance aa') is 2 inches,
present the following design data:
(a) A sketch to scale of the nozzle contours.
(b) The height of the test section, dd', and the width of the tunnel a t the
throat and a t the test section.
(c) A plot of pressure versus distance along the axis of the tunnel, and
along the wall contour.
(d) The maximum half-angle of divergence of the expanding portion of the
nozzle.
(e) Draw the streamline which a t the throat is midway between the axis
and wall. Compare the length-height ratio of a nozzle using this
streamline as a wall contour with the length-height ratio of the sharp-
corner nozzle.
15.24. Determine the form of the hodograph characteristics for the tangent
gas of Chapter 11.
16.26. Show, for adiabatic flow, that a plot with polar coordinates M* and
a traces out an ellipse.
Demonstrate how an M*-or ellipse, drawn on a sheet of tracing paper to
the same scale as the M * 4 characteristic curves, may be used in conjunction
with the latter for graphically determining wave directions without drawing
normals to the hodograph characteristics.
Chapter 16
OBLIQUE SHOCKS
NOMENCLATURE-Continued
Streamline
control/
Svrfgre
FIG.16.2. Nomenclature for oblique-shock analysis.
lines forming part of the control surface are exact duplicates of each
other, we may write
Continuity.
P I J ' ~ ~= PZV~Z
Momentum in t-direction.
(PIVnl)Vtl= ( P Z V ~ ~ ) V-'.~Vt2
Z ;= Vtl = Vt (16.2)
Momentum in n-direction.
2
PI - ~2 = P Z V ~-
Z~ 1 V n l ~
Energy.
cP(T1- T2) = (VZ2- V I 2 ) / 2
The absence of change in the tangential component of velocity
(Eq. 16.2) bears out the change-in-coordinate point of view of Art. 16.1.
From Eq. 16.2 and the geometry of the velocity triangles,
Eqs. 16.1, 16.2, 16.3, and 16.5, together with the perfect gas relations,
define completely the relations between the states on the two sides of
Art. 16.2 OBLIQUE SHOCK EQUATIONS 533
the shock. For example, suppose that state 1 is completely specified hg-
the values of pl, pl, Vnl, and Vtl. Then the four equations mentioned
are suffcient to determine p2, p2, Vn2, and Vtz. Since Vnl and Vtl to-
gether specify the shock angle u, and since Vn2and Vt2together specify
the angle (a - a), the solutions also yield the shock angle u and the
turning angle 6. Furthermore, the values of p and p define the sound
speed c, so that the Mach Numbers fore and aft of the shock, MI and M2
respectively, may also be found.
The basic shock relations of Eqs. 16.1: 16.2, 16.3, and 16.5 may be
combined in many ways to yield a great many algebraic relations relat-
ing various shock parameters.'', 3 , Practically, we wish to obtain
2n
P2 v n 2
= plvn12(l- );
from which
Similarly, we may solve for Vnz from Eqs. 16.3 and 16.1,
Substituting Eqs. 16.6a and 16.6b into Eq. 16.5, and rearranging, we
obtain the prcssurc ratio as a unique function of the density ratio, or the
density ratio as a unique function of pressure ratio:
534 OBLIQUE SHOCKS Ch. 16
k - 1 ~ 2
+ Vnz2+2 Vtz2 - k-5 POP O
Introducing Eqs. 16.1 and 16.2, and reducing algebraically, we finally get
If we had factored out p2V,22 in Ea. 16.3, we would have arrived in-
stead a t
Eqs. 16.8, 16.13, 16.14, and 16.15 constitute four relations among the
variables u, 6, MI, M2, p2/pl, and p21P1. Any two of these variables,
therefore, may be taker, rts independent parameters, the values of which
completely determine the values of the four remaining variables. The
temperature ratio, entropy change, and ratio of isentropic stagnation
pressures may of course be calculated readily after the variables named
above are specified.
-
Deflection
M,, I n i t i a l M a c h Number
FIG.16.5. Exit Mach Number versus inlet Mach Number, with turning angle as
parameter (k = 1.4).
and po,/poZ of the table are respectively the values of p2/pI, P2/Pl, T2/T1,
and p~,/pop,,for an oblique shock with inlet Mach Number M1 and shock
angle u. From the geometry of Fig. 16.2 it follows further that M2 =
M,/sin (a - 6).
16.3. Shock Geometry
Application of Second Law of Thermodynamics. The shock formulas
as previously developed do not distinguish between a compression dis-
continuity (pressure increase) and a rarefaction discontinuity (pressure
decrease). However, since the second law of thermodynamics allows
only entropy increases for adiabatic processes, it follows that s2 2 sl.
Remembering that an oblique shock may be reduced to a normal shock
by means of a coordinate transformation not involving entropy changes,
and that in a normal discontinuity the entropy decreases for a rarefac-
tion shock, we conclude that the only type of oblique discontinuity
which is allowable is one in which there is a pressure rise.
BENDING OF STREAMLINE. Since p2 > pl, it necessarily follows from
the continuity relation (Eq. 16.1) and the momentum theorem (Eq.
16.3) that Vnl > Vn2. Then, since Eq. 16.2 requires that Vtl = Vt2,
the velocity triangles of Fig. 16.2 show that 6 must always be positive
in magnitude. In other words, we may state the general rule that in
crossing a n oblique shock the streamline always bends toward the shock line.
DECAY OF RAREFACTION SHOCK AT CONVEX CORNER. The impos-
sibility of a rarefaction shock is of course also connected with the fact
Dl/ / Rarefaction /
/ / /
Streamline ,'/ / /n2
/
/
Possible
that rarefaction waves always tend to become less steep. Thus, in the
expansive flow around a corner, Fig. 16.7, there is a rarefaction discon-
tinuity confined to the singular point of the sharp corner, but this dis-
continuity rapidly decays to a continuous Prandtl-Meyer expansion
which fills the entire expansion region.
540 OBLIQUE SHOCKS Ch. 16
/
/
Impossible
find that each streamline reverses itself twice a t cusps. This means that
a t each point in the compression region there would exist simultaneously
three different pressures and velocities! Mathematically this is con-
ceivable with the aid of multi-sheeted surfaces branched a t the Mach
lines, but it is physically impossible. The oblique compression shock
provides a means for carrying out a sudden change in direction without
involving this difficulty. The reversal of the hypothetical Prandtl-
Meyer compressive flow occurs a t cusps whose loci are known as limit
lines. This example, and the example of flow reversal in a source-type
flow a t a limit line where M = 1,
are perhaps the simplest illustra-
tions of the limit line phenomena
/ which are discussed more fully in
Volume 11, Chapter 20.
Relation of Shock Line to Mach
L i e s . In Fig. 16.8 the shock line
is shown as lying between the
Mach lines of the upstream and
downstream flows. We shall now
demonstrate that this is always
FIG. 16.9. Illustrates that left-running SO.
shock "overtakes" upstream left-running Referring to Fig. 16.9, let 11,
Mach wave, and is overtaken by down-
stream left-runniner- Mach wave. and 112 represent the respective
directions of the left-running Mach
lines in regions 1 and 2, making the angles a, and a2 to the upstream
and downstream flows, respectively.
Art. 16.3 SHOCK GEOMETRY 541
which demonstrates that the shock line i s more steeply inclined to the $ow
than a Mach wave in region 1.
SHOCK IS OVERTAKEN BY DOWNSTREAM M ACH LINES. We note from
Eq. 16.12 and the definition of cuz that
/
/-Shock Line
(a) (b)
FIG. 16.11. Hodograph shock polar.
(a) Configuration of velocity vectors.
(b) Construction for determining shock angle and final velocity.
required to lay off the vector V 2 as well as the shock angle (T may be
found from the preceding formulas, and thus point 2 may be located in
the hodograph plane. The upstream and downstream velocity com-
ponents along and parallel to the shock line must be related t o each
other as shown in Fig. 16.11a, with VL1= V t 2= V t . It is to be noted
that Fig. 16.11a may be found by merely superposing the velocity tri-
angles of Fig. 16.2.
Now, for each value of Vl/c* there are an infinite number of end states
(corresponding to different values of 6) in the hodograph plane which
can be reached from state 1. The locus of these points is known as the
hodograph shock polar for MI*. A typical shock polar is shown in Fig.
16.11b. It is symmetrical about the horizontal axis, since there are two
possible shock waves for given values of M 1 and 6, one running leftward
relative t o the flow, and the other rightward. Having once constructed
the shock polar for a certain value of V l / c * , the solution for a given
turning angle 6 is found by laying off from the origin a line a t the angle 6.
Where this line intersects the shock polar is the end point of the velocity
vector Vz/c*. The normal OA to the extension of the straight line ?-2
then gives the direction of the shock line.
Art. 16.4 SHOCK POLARS 543
This formula allows us to plot v2/c* versus u2/c* for any specified value
of the parameter ul/c*.
544 OBLIQUE SHOCKS Ch. 16
-
pE,; 2w
for M , -
o 6
FIG. 16.14. Illustrates strong and weak shocks.
in a duct of constant area. For this case the turning angle 6 is zero, and
the strong solution on the shock polar of Fig. 16.15 is a normal shock,
while the weak solution is an infinitesimal wave of zero pressure rise.
From our previous discussions of one-dimensional flow in ducts, i t is
clear that the solution which occurs in practice depends on the boun-
dary conditions, in particular on the ratio of downstream pressure to
upstream pressure. If the downstream pressure is sufficiently low, thc
weak solution occurs, which is to say that there are no shocks a t all.
I
(Normal (Plane
Shock) Sound Wave)
FIG.16.15. Strong and weak shocks for 8 = 0.
graph by stating that whether the oblique shock is strong or weak de-
pends, a t least in part, on the boundary conditions as to pressure. If
the downstream pressure for the two-dimensional flow, Fig. 16.16a, is
sufficiently small, a weak shock will occur a t corner A and the flow
approaching corner B will usually be supersonic. If the back pressure
is sufficiently high, on the other hand, the shock emanating from corner
0 @ Subsonic H
A will be of the strong type, and the flew approaching corner B will
usually be subsonic.
As will be seen later, however, an oblique strong shock cannot extend
to a wall, but rather curves as shown in Fig. 16.16b until it is a normal
shock at the upper wall. Each streamline therefore passes through a
shock of different strength, and each point on the shock line is repre-
sented by a corresponding point on the segments between 2s and 4 of
the shock polars of Figs. 16.16~and 16.16d. The flow behind the shock
is subsonic, the streamlines are curved, and there is vorticity or rotation
in the flow downstream of the shock because of entropy variations from
streamline to streamline.
When a symmetrical wedge of half-angle 6 is moved through the
atmosphere at supersonic speed, only the weak shock is observed in
practice. This may be rationalized in the following way. Suppose that
Art. 16.5 SPECIAL ASPECTS OF OBLIQUE SHOCKS 547
the upper wall of Fig. 16.16a is moved infinitely far upward. Then in
the limit, we obtain the upper half of the flow pattern for the wedge
moving a t supersonic speed. In the atmosphere, however, the pressures
far downstream of the wedge can differ from those far upstream by only
infinitesimal amounts, and hence the boundary conditions are such that
the weak shock occurs. Thus, with supersonic aircraft, the obliquc
shocks seem to be invariably of the weak type, except under conditions
described below where the geometry of the aircraft does not allow of
attached shocks.
The schlieren photographs of Fig. 16.17 support the view that the
boundary conditions determine whether the weak or strong shock a p
pears in practice. All three photographs show the flow near the entrance
,5 10 20 30 40 50
8 ( o r 81, Degrees
Fro. 16.18. Maximum turning angle For a given initial Mach Number.
(a) Pressure-deflection shock polar.
(b) Hodograph shock polar.
(c) Chart for k = 1.4.
it is evident that a t this limiting condition the weak and strong oblique
shocks become identical. The relation between MI and 6,, or between
MI, ,in and 6, is shown in Fig. 16.18~. No solutions may be obtained
in the shaded area below the curve in the graph. (See Problem 16.11a.)
Consider the case of uniform, parallel flow a t Mach Number M1 past
a wedge of half-angle 6. If 6 is less than 6,, the shock is attached to
the wedge (Figs. 16.19a and 14.8). However, if 6 is greater than 6,,
the shock cannot be attached to the wedge, for this would require the
streamline approaching the point of the wedge to turn through an angle
greater than 6,. Under these circumstances, we observe in practice
that the shock is detached from the wedge (Figs. 16.19b and d). Simi-
larly, a detached shock always stands in front of a blunt body (Fig.
16.19~).
Art. 16.5 SPECIAL ASPECTS OF OBLIQUE SHOCKS 549
to the given MI. Behind the detached shock the flow'is in part super-
sonic and in part subsonic, leading to great difficulties in analysis be-
cause of the radically different properties of the differential equations
for subsonic and supersonic flow.
Similar considerations apply to the flow past a concave corner in a
duct (Fig. 16.20).
, ,a
FIG. 16.21. Relation between and A*.
entire range of Mach Numbers from unity to infinity. This means that,
except in a very narrow range, the flow behind a weak shock i s supersonic
and the flow behind a strong shock i s subsonic. (See Problem 16.11b.)
Nearly Normal Shocks. From the geometry of the hodograph shock
polar, Fig. 16.11b, it may be seen that a strong shock of small turning
angle will have a shock angle a of nearly 90'. From the 6--p shock polar
of Fig. 16.13, it may be seen furthermore that the pressure rise across
such a shock is substantially the same as that across a normal shock.
Art. 16.6 VERY WEAK SHOCKS 55 1
Thus we have the approximate rule that strong shocks of small turning
angle are nearly equivalent to normal shocks.
From this we see that as Ap goes to zero, Ap also goes t o zero. The
limiting relation between Ap and Ap, however, as Ap/pl goes to zero, is
evidently
AP/AP = k ~ l / =
~ lc12
Thus we see that for weak shocks, Ap/Ap = c12. But, since c2 =
(ap/ap),, it follows that the changes in pressure apd density in a very
weak shock are connected in the limit through the isentropic relation
(see Fig. 16.3).
Combining this result with Eq. 16.7, an'd noting that, for very weak
shocks, Eqs. 16.1 and 16.3 predict that 17,~approaches Vnl, we get
.
Shock
Polor
\
\
FIG. 16.22. Normal to hodograph FIG. 16.23. Hodograph shock polar and
shock polar at 1 has same direction as hodograph characteristic p'assing through
Mach line. point 1 have equal slopes and radii of
curvature.
Expansion ( 8 )
FIG. 16.24. Reflection of shock from constant-pressure boundary (see also Fig.
14.21).
(a) Hodograph diagram.
(b) Streamlines and waves.
Shock P o l a r
.<
I
3 "/c*
j \shock polar
For S t a t e 2
The graphical solution to this problem with the aid of the shock polar
diagrams is shown in Fig. 16.27b and 16.27~. The latter diagram is
especially convenient for this type of problem.
Since fields 4' and 4" have the same pressure and same stagnation
temperature, the adiabatic and isentropic relations require that the
field with the lower s t a g n a t i o a y (higher entropy) have the lower
(c) (4
Fro. 16.27. Crossing of shock waves.
(a) Streamlines, shock lines, and slip line.
(b) Hodograph diagram.
(c) Pressure-deflection diagram.
(d) Schi~erenphotograph showing refraction of two shocks of equal strength;
may also be interpreted as reflection of shock from straight boundary. (Re-
produced by permission of The Macmillan Co. from Elements of Aerodynam-
ics of Supersonic Flows, by A. Ferri, 1949.)
velocity. In this case shocks A and C are stronger than B and D, hence
the gas speed is larger in field 4" than in field 4'. This same rule may be
applied whenever vorticity appears in an adiabatic flow.
If shocks A and B are of equal strength, i.e., if 62 = 83, then the shock
configuration will be symmetrical, and no slip line will appear (Fig.
16.27d). The streamline passing through the junction point will undergo
no change in direction, and the entire process will therefore be equivalent
to that of Fig. 16.25, that is, to shock reflection from a plane wall.
The type of regular interaction shown in Fig. 16.27 may prove to be
impossible if MI is too small or if 62 and 63 are too large, for the same
reasons that regular reflection may be impossible with the incident wave
pattern of Fig. 16.25. A Mach reflection then occurs.
Art. 16.7 REFLECTION AND INTERACTION OF SHOCKS 557
Mach Reflection. When MI is so small or 6 is so large as to precIude
the possibility of regular reflection in Fig. 16.25, a Mach reJEection (Fig.
16.28) occurs. Near the upper wall the shock must necessarily be nor-
mal, for this is the only way in which a shock can occur without a change
in direction. Shock C, of opposite family to shock A , emanates from
lrrotationol
Flow Rotational
out that this approximation very often gives acceptable results. I n es-
sence, the approximate method ignores the variations in entropy change
across different parts of the oblique shock, and, therefore, the slip lines
and the waves reflected from the shock do not appear. Fig. 16.30~ shows
the simplification in wave pattern afforded by this approximation.
I n the region of vorticity where continuous waves are present, the
entropy represents an additional variable. The method of character-
istics may be worked out formally in this case, as in Art. 15.12.
Without formally working out the method of characteristics for non-
isentropic flow, however, we may make calculations for the region of
rotational flow by extending the ideas already presented. In isentropic
flow the pressure is a unique function of velocity, whereas in nonisen-
tropic flow this is not true. This difference explains why the hodograph
characteristics are fixed for two-dimensional, isentropic flow but are not
fixed for nonisentropic flow. These remarks suggest further that we
apply the method of characteristics stepwise to the rotational flow pat-
terns, using not the hodograph characteristics diagram, but rather work-
ing out the physical characteristics in terms of the relations between
pressure change and direction change for each individual wavelet, taking
note of the actual pressures, velocities, and Mach Numbers in each field.
This method is in essence equivalent to the more formal mathematical
methods.
SHOCKS CROSSING SLIP LINES. When a shock enters a region of
parallel flow with varying speed, it is refracted and suffers partial re-
flection. This problem is discussed in Reference 14 (see Problem 16.36),
and, more generally, in Reference 18.
2 C 3 ( ~ e )+
~ c(ae), + C 2 ( ~ e )=t
P - Pw
1
= 3 ... (16.20)
zkpmMm2
where A0 is positive when counterclockwise. The upper and lower signs
refer respectively to left-running and right-running Mach waves, and
the coefficients are as follows:
(16.21~)
It will be noted that the first term of Eq. 16.20 represents the linear-
ized solution.
Series Solution for Oblique Shock. If the shock relations of Art. 16.2
are expanded in a similar way, letting subscript 1 refer to conditions
upstream of the shock, we find (ll.*) that the first two terms are the
same as for isentropic waves of one family, with the difference between
the two types of flow appearing initially in the third term:
where
Although the coefficients C1, Cz, C3, and D appear complex, they
depend only on M, and k, and need be computed only once. The coeffi-
cients C1 and Cz are tabulated in Table B.8 and are plotted in Fig. 16.41.
Art. 16.10 EXAMPLES OF FLOWS CONTAINING SHOCKS 563
Equations 16.20 and 16.22 may then be uscd for obtaining analytical
results, provided of course that waves of opposite family do not cross
each other. Examples are given later.
If only terms up to second order in A0 are employed, it is clear from
these equations that the oblique shock theory and the continuous wave
theory give identical results.
are extremely weak and the velocity difference across the slip lines i s
OBLIQTJE SHOCKS Ch. 16
nature of the flow pattern for such a jet when the nozzle is of the con-
verging type and the flow is underexpanded. The supply pressure is
assumed to be 100 psia, and the exhaust region pressure 27 psia. Hence
sonic flow exists a t the throat with a pressure of 53 psia. The wave
pattern is worked out for a 2" turning angle across each wave, except
for the shock. Entropy increases across the latter are ignored. Expan-
sion wavelets are shown by dashed lines, and compression wavelets by
solid lines.
An important result demonstrated by Fig. 16.33 is that the average
jet direction differs significantly from the nozzle direction because of
the general bending of the jet. This bending effect reduces the work
available from the turbine wheel.
Having found the pressure distribution over the profile, the resulting
forces and moments may be calculated, leading finally to the lift, drag,
and moment coefficients. Although the "exact" theory as given here is
more accurate than the linear and second-order theories, it has the dis-
advantage that a separate set of numerical calculations is required for
each combination of profile shape, angle of attack, and Mach Number,
and that it is not possible to arrive a t simple analytic formulas showing
the effects of these variables.
Eggers and Syvertson ( ' 9 ) have formulated a comparatively simple
approximation to the shock-expansion method for thin profiles, which,
for Mach Numbers greater than 3 and flow deflection angles less than
25", yields pressure distributions in error by less than 10yo.
In order to determine the nature of the wake immediately behind the
profile, we note that the two shocks emanating from D must be of such
strengths that the pressures and flow directions in regions 6 and 7 are
the same. The two boundary streamlines passing through D have
passed through shocks of different strength, and, consequently, have
different stagnation pressures. Accordingly, the double streamline trail-
ing from point D is a slip line, or vortex sheet. The common flow direc-
tion of fields 6 and 7 is in general different from that in region 1, and,
furthermore,'the common pressure in these fields differs from that of
region 1. These differences, however, areusually minute.
The "exact" calculation is very much more difficult if the waves
generated a t the profile interfere with each other so near the profile
that the reflected waves reach the profile further downstream, as is the
case with the circular-arc profile of Fig. 16.36. Consider the flow over
the upper surface (Fig. 16.3Ga). An oblique shock is generated a t the
nose in accordance with the turning angle associated with the tangent,
at the nose. Because of the continuous curvature of the profile, however,
Art. 16.11 TWO-DIMENSIONAL PROFILES 569
FIG. 16.38. Experimental drag polars FIG. 16.39. Experimental drag polars
for strut (after Busemann and Walchner). for subsonic profile (after Busemann and
Walchner).
chord line; similarly the ordinate V L of the lower surface is taken posi-
tive when below the chord line. The angle of attack, a, is positive when
directed as shown in the sketch.
The method of analysis is similar to that of the linearized theory
(Art. 14.6), but the second-order terms of Eqs. 16.20 and 16.22 are
taken into account. During the course of the calculations several inte-
grals appear which depend only on the geometry of the profile. These
are
Only pressure drag is included in the drag coefficient given above. The
skin-friction drag coefficient must therefore be added to the right-hand
side of Eq. 16.26 in order to get the total drag coefficient.
The moment coefficient of Eq. 16.27 is reckoned about the leading
edge, and is taken to be positive when the moment acting on the profile
-
"m E Chord-' E
F IG. 16.42. Cambered single-wedge
profile.
(c)
state the general rule that the results of the second-order and linear
theories approach each other as the profile shape approaches a flat plate.
T RIANGULAR PROFILE. TO illustrate the difference between the
second-ordcr and linear theories, consider the triangular profile of Fig.
16.42 a t zero angle of attack. The various J-integrals are
The first-order theory predicts zero lift, but the second-order theory
predicts a negative lift. The latter fact is especially interesting in that
it shows still another surprising difference between subsonic and super-
sonic flow. In subsonic flow at zero angle of attack with a positive cam-
ber as in Fig. 16.42, the lift is also positive; in supersonic flow the lift
for the same profile is negative.
At zero angle of attack, the thickness drag for this profile is the same
for both the first-order and second-order theories.
If lc.p, is the distance from the leading edge to the center of pressure,
we may show by taking moments that
Thus, the second-order theory predicts that the c.p. is a finite distance
forward of the mid-chord point, whereas the linear theory predicts it
to be infinitely forward. As a general rule the most significant improve-
ment of the second-order theory over the linear theory lies in a more
accurate prediction of the center of pressure.
COMPLETELY SYMMETRICAL PROFILE. If the profile is symmetrical
both laterally and longitudinally, we may set Jr. = J uand we may also
set J2= 0. We then obtain
CL = 2Cla (16.28a)
+ 2C1Jlu
CD = 2Cla2 (16.28b)
Cm = Cla + 4C2J3ua (16.28~)
No second-order terms appear in the expressions for CL and CD. We
may write, for the location of the center of pressure,
Art. 16.11 TWO-DIMENSIONAL PROFILES 575
0 Ferri's Experiments
- "Exact" Theory
-.- Second O r d e r Theory
--- Linearized Theory
% Chord
FIG. 16.44. Comparison of experimental pressure distribution with results of exact,
second-order, and linear theories (after Ferri).
Fig. 16.45 shows that the second-order and exact theories give virtu-
ally identical results for lift and drag up to angles of attack of the order
of 10". Both provide a significant improvement over the linear theory,
and compare very well with the experimental data. The agreement
with the moment data is only fair. It is a t first surprising that the
actual drag is slightly less than the theoretical pressure drag of the exact
theory, since the latter does not include skin friction. This effect is
associated with the shock-boundary layer interaction near the trailing
edge (Art. 16.12).
Downwash from Thin Profile. The flow aft of a profile is of interest
if other bodies should have to be placed in the wake. For instance,
Art. 16.11 TWO-DIMENSIONAL PROFILES
Theory Includes
a, Degrees
6.3% Thick
-
-.-
0 Ferri's Experiments
Exact Theory
--- Second O r d e r T h e o r y
Linear Theory
control surfaces usually are in the wake of the main lifting surface of an
aircraft.
If not for entropy increases across the shock waves (Fig. 16.46), there
would be no slip line between zones 4 and 5 and the upwash angle @
would be zero. It is therefore necessary to include terms of third order
in Eqs. 16.20 and 16.22 in an investigation of the upwash angle. Such
an analysis (lo' yields the following results for a flat-plate airfoil when
578 OBLIQUE SHOCKS Ch. 16
These results were found to check within 1% of the exact theory for
a = 10' and MI = 5. The magnitude of 0 is generally small, being al-
ways less than 0.06' for a = 5.7' over the range of M, from 1.2 to 5.0.
A surprising feature is that, except for M, less than about 1.3, there is
an upwash behind the wing, rather than a downwash as would be ex-
pected for subsonic flow. Of course momentum considerations require
that a lift on the profile must be accompanied by a net downward change
of momentum flux of all the air influenced by the profile; hence if the
part of the airstream near the profile suffers an upwash, other parts of
the airstream must acquire a downwash.
F I ~16.46.
. Nomenclature for upwssh relations.
(a) Flabplate profile.
(b) Doubly symmetric profile.
When (i) the profile has thickness, (ii) there are four oblique shocks a t
the nose and tail (Fig. 16.46b), and (iii) the airfoil is completely sym-
metrical, it may be shown ( l o ) that
This formula indicates that the flow deflection for a profile with thick-
ness is considerably larger than for a flat plate. For example, when
E = a, the ratio between the two is 16.
very large distances from the profile the flow is, for all practical purposes,
again uniform and parallel, with only infinitesimal deviations from the
free-stream condition and with all waves a t the free-stream Mach angle.
A large part of the wave system for the lower side of the profile is omitted
from Fig. 16.47 so as not to confuse the pattern unnecessarily.
It is this type of reasoning which justifies the analysis accompanying
Fig. 16.37.
FIG. 16.47. Wave pattern downstream of flat-plate airfoil. Most of the wave
system on the underside is omitted in the interest of clarity.
(a) (b)
FIG. 16.48. Busemann biplane.
(a) Zero thickness drag.
(b) Small thickness drag.
Even when shocks are present, as in the biplane design of Fig. 16.4813,
they are confined to the space between the profiles and hence produce
only a fraction of the drag which they would otherwise produce if allowed
to propagate out to great distances from the profile.
(tj) (c)
FIG.16.49. Reflection of oblique shock from boundary layer on flat plate.
(a) Schematic representation.
(b) Schlieren photograph for turbulent boundary layer (after Liepmann, NACA).
(c) Schlieren photograph for laminar boundary layer (after Liepmann, NACA).
such a way that compression wavelets are generated along the concave
part of the boundary layer HJ. These wavelets coalesce into the shock
KL.
Experimental observations are in qualitative accord with the fore-
going rationalization of the wave pattern. The nature of the interac-
tion depends upon Mach Number, shock strength, velocity profile in
the boundary layer, and especially on whether the boundary layer is
laminar or turbulent. In all cases observed, however, the flow patterns
contain in greater or lesser degree all or part of the elements shown in
Fig. 16.49a (see also Chapter 28 of Volume 11).
Art. 16.12 SHOCK-BOUNDARY LAYER INTERACTION 583
normal shock could not possibly extend to the wall when a boundary
layer is present and that the rapid thickening of the boundary layer up-
stream of the shock gives rise to an oblique shock. The reflected shock
a t the juncture of the normal and oblique shocks is necessary as a result
of continuity requirements. Thus, a series of forked normal shocks ap-
pears, with the length of the normal shock becoming progressively
smaller as the boundary layer thickens. Finally the sonic velocity is
reached and the stream then decelerates subsonically as it refills the'
passage. In place of a normal shock, therefore, the stream contracts by
means of a combination of normal
and oblique shocks to Mach Num-
ber unity and then diverges to
subsonic speeds.
Shock Propagated at a Bend.
When no boundary layer is present
a shock is propagated in a straight
line from a sharp concave bend, as
in Fig. 16.1~.
The corresponding pattern for
supersonic flow past a sharp con-
cave corner with a FIG. 16.51. Generation of shock in con-
layer cave corner with boundary layer present.
present is shown in Fig. 16.51.
The large pressure rise near the corner propagates upstream through the
subsonic boundary layer. As the boundary layer decelerates, it also
thickens, and thus continuously propagates compression wavelets into
the supersonic flow. These wavelets merge and ultimately form a shock
across which the turning of the stream occurs.
584 OBLIQUE SHOCKS Ch. 16
FIG.16.52. 1 % ~
rwar trailing edge of supersonic profile.
(:I)Boundary layer abscnt.
(b) Boundary layer prescrlt.
16.11. (a) Show that the shock angle urn, corresponding to the maximum
turning angle 6, is given explicitly in terms of MI by the following expression:
(b) Show that the shock angle u* corresponding to the turning angle 6* for
which the downstream Mach Number is unity is given explicitly in terms of MI
Ch. 16 PROBLEMS
by the following expression:
16.13. (a) Demonstrate that if two points in the flow have the same pressure
and same stagnation temperature, but different stagnation pressures, the veloci-
ties are related as follows:
(b) From the above result show with the help of momentum considerations
that the pressure drag of a two-dimensional supersonic profile in an infinite
medium is directly related to the entropy increases across the shock waves in
the flow field.
(c) From the result in part (a) show that when a slip stream appears in an
adiabatic flow, the Mach Number is higher on the side of the slip stream having
the lower entropy.
(d) Show that if the oblique shocks produced by a supersonic, two-dimen-
sional profile did not interact with rarefaction waves in such a way that the
oblique shocks were ultimately weakened to become Mach waves, the drag of
such a profile would be infinite.
16.14. A parallel stream with a pressure of 10 psia and a Mach Number of
2.0 approaches a sharp l o 0 corner (see sketch).
(a) Assuming that the shock is of the weak variety, calculate the following
quantities: the final Mach Number, the final pressure, the change of entropy,
and the ratio of the final flow per unit area to the initial flow per unit area.
(b) Make a sketch to scale showing the streamlines, the shock line, and the
Mach lines upstream and downstream of the shock.
(c) Compare the results of part (a) for final Mach Number and final pressure
with the analogous result for a 10"-turn based on linearized theory, a 10"-turn
based on the Prandtl-Meyer corner-type flow, and a lo 0 -turn based on the
second-order theory.
588 OBLIQUE SHOCKS Ch. 16
16.15. Compare the linearized theory with the oblique-shock theory and the
second-order theory by plotting curves of final Mach Number and of the frac-
tional change in pressure (Ap/p) against the deflection angle, basing each curve
on a constant initial Mach Number.
Use constant initial Mach Numbers of 1.5 and 2.5, and carry each curve from
zero deflection to the maximum deflection angle.
16.15. Air flows in the passage shown in the sketch with an initial Mach
Number of 2.0. Determine the maximum turning angle 6 for which three regular
reflections of the original shock arc= possible.
ratio p,/p,, for values of the latter between unity and the value for which a
normal shock stands in the exit plane. (b) Determine the range of pressure
ratios, p,/p,, for which regular reflection of the oblique shock wave originating
a t the nozzle exit is not possible.
Symmetrical
t Porabola
........ ....................................
..~ :..:..:
............... ~ . .~,
:.i
~ . . :....... ...............:
... ..........
.......;c..
:>:.:::::
:<.<x...:.:.,.,.
?.................... ...., ..................
.:.::::..?:,~5*
............................................
I.."
16.34. Using second-order theory, find Cr.and Cn for the profile shown in
the sketch in terms of &, t/1, and a.
692 OBLIQUE SHOCKS Ch. 16
16.35. A two-dimensional double-wedge profile (see sketch) is at zero angle
of attack in an air stream of Mach Kumber 2.016. Ignoring friction, and using
exact methods,
(a) Calculate the lift coefficient.
(b) Calculate the drag coefficient.
(c) Calculate the ratio Lc.,,/l where I,., is distance aft of the leading edge
to the center of pressure, and I is the chord.
(d) Estimate the downwash angle, E , immediately downstream of the trailing
edge.
. Sketch the wave system s t great distances from the profile.
it: :ompare the results obtained above with similar results based on the sec-
ond-order theory.
16.36. (a) Demonstrate the following relation between turning angle, initial
Mach Number, and shock strength exp~essedin terms of the fractional pressure
k+lpz-pi
- ( 1 +-Ui )
tan 6 = rt -1 P 2 - PI
k PI k+lpz-pi
Pl
(b) From this show that, if an obliyw -hock crosses a slip line which divides
parallel flows having Mach Numbers Mia and MI=, the condition that the shock
be transmitted without any reflection whatsoever is that
~1.4' (1 + k + 1 7 ) ~PI Pi )
-
-
P2 - Pi ( 2
PI
+ P2 - Pl)
pl
MIB2 =
(c) Show from part (a) that for a fixed pressure ratio across the shock, there
is a maximum turning angle, 6, (,I, which occurs a t a Mach Number given by
(d) From part (c) show that if an oblique shock crosses a weak slip line divid-
ing parallel flows of nearly the same Mach Number, the condition that there
shall be no reflection a t the slip line is approximately
APPENDIX
Appendix A
THEORY OF CHARACTERISTICS
A S . Introductory Remarks
The differential equations describing the motion of compressible
fluids in two or three dimensions are, with the exceptions of a few simple
cases or under greatly simplifying assumptions, not easily soluble by
analytical methods. Fortunately, there is a class of problems for which
solutions may be rather easily obtained by means of numerical or
graphical procedures. Such is the case when the motion is described by
a partial differential equation of hyperbolic type. We shall see that
the possibility of constructing solutions arises from the special properties
of certain characteristic curves which satisfy second-order, partial differ-
ential equations of hyperbolic type with two independent variables.
The purpose of this Appendix is to outline, for those readers not al-
ready familiar with the subject, the mathematical theory underlying
the method of characteristics. Particular attention will be given to
those aspects of the theory which are relevant to problems of compress-
ible flow (readers are referred to Reference 1 for a more complete treat-
ment than is given here).
NOMENCLATURE
functions of x, y, q,, and velocity potential
Pu
speed of sound signifies conditions along
ratio of specific heats characteristic curve
radius in cylindrical and signifies conditions along
polar coordinates characteristic of family
radius in spherical coor- 1.
dinates signifies conditions along
time characteristic of fam-
velocity components ily I I .
Cartesian coordinates
neither viscosity nor thermal conductivity, and that there are no dis-
continuities in fluid properties such as occur in shock waves):
Two-Dimensional, Plane, Supersonic, Irrotational, Steady Flow
(Chapter 16). Application of Euler's equation, the continuity equation,
and the condition of irrotationality yields
(c2 -
2
9z )9m + (c2 - '~Y2)~yy= 0
- 29z(~y9xy (A.1)
where cp is the velocity potential, x and y are the Cartesian space coordi-
nates, c is the local sound velocity, and fixdenotes d29/axz, etc. The
velocity components u and v are given by
The local sound velocity depends on 9=and cpu according to the energy
equation,
k-1
c2 = co2 - -(9x2
2
vU2)+ (A.2b)
and c is the local sound velocity. The fluid velocity in the x-direction
is given by
u = cox (A. 6a)
The local sound velocity is related to P
(, and cot through the equation
(A.6b)
I 0 dcp, dy I-
-
+
A dy dcp, - D dy dx C dx dp,
(A.14)
A dy2 - 2B dx dy f C dx2
At any point (x, y, cp) on the integral surface, the values of rp, and p,
are known from the shape of the surface. Consequently the coefficients
A , B, C, and D are known a t all points of the surface. I n general, there-
fore, the value of rp,, (and, by similar reasoning, the values of p,, and
p,,) is determined by Eq. A.14, except when both the numerator and
denominator of Eq. A.14 are equal t o zero. The latter condition gives
the answer to our original query, for when pz,, p,,, and p,, are indeter-
minate, they may be discontinuous.
Physical Characteristic Curves. By setting the denominator of Eq.
A.14 equal to zero, we obtain the differential equation of the projections
of the characteristic curves in the physical plane, namely,
(A. 15a)
B f d B 2 . - AC
(A. 16)
c
Eq. A.16 gives the slopes of the characteristics in the pX,9,-plane in
terms of q,, 9,,x, and y. It may be recalled that Eq. A.15b gives the slopes
of the characteristics in the x, y-plane also in terms of qz, 9,, x, and y.
Given suitable initial data, therefore, the characteristic curves in the
physical and hodograph planes may simultaneously be constructed nu-
merically or graphically in stepwise fashion. Since the physical data of
interest (such as pressure, velocity, etc.) may be found from the values
of qz and 9, a t each point, a solution of the problem is thereby deter-
mined. Before going into details, we shall discuss certain general fea-
tures of the method.
Classification of Differential Equations. Examination of Eqs. A .15
and A.16 shows that the differential equation A.ll may be of three
types, depending on the sign of (B2 - AC). These types are defined as
follows:
Hyperbolic type. The value of (B2 - AC) is positive, so that Eqs.
A.15 and A.16 each have two real roots. Two characteristic curves pass
through every point of the physical and hodograph planes.
Parabolic type. The value of (B2 - AC) is zero, and there is one real
root to Eqs. A.15 and A.16. This type is not of great practical signif-
cance.
Elliptic type. The value of (B2 - AC) is negative, so that Eqs. A.15
snd A.16 have no real roots, and the characteristic curves are imaginary.
Applying these criteria to Eqs. A.l, A.3, A.5, A.7, and A.9, it is found
Lhat (i) for the two cases of steady motion the differential equation is
hyperbolic for supersonic flow and elliptic for subsonic flow, whereas
(ii) for the three cases of non-steady motion the differential equation is
nlways of hyperbolic type. Hence real characteristics always exist for
the non-steady motions, but exist for the steady motions only when the
flow is supersonic.
APP. A THE CHARACTERISTIC CURVES 601
X
(4
FIG.A.1. Characteristics net.
(a) Physical plane.
(b) Hodograph plane.
known a t each point of the physical plane and consequently the entire
flow pattern in terms of the velocity distribution is established. In the
stepwise construction the curvilinear characteristics are replaced by
straight-line segments connecting the lattice points of the network.
To establish a point such as 7 and its hodograph image 7', we proceed
as follows:
(i) Using Eq. A.l7a, the value of (dyldx)~a t point 4 is computed,
using the values of A, B, and C corresponding to point 4. A line with
the slope (dyldx)~is tentatively passed through point 4.
(ii) Using Eq. A.l7b, the value of ( d y l d z ) ~a ~
t point 1 is computed,
using the values of A, B and C corresponding to point 1. A line with the
slope ( d y l d x ) ~is~tentatively passed through point 1. Its intersection
with the line drawn in step (i) establishes, a t least approximately, the
location of point 7.
(iii) The value of (dqg/dcpz)r at point 4' is computed from Eq. A.18a,
and a line is tentatively drawn through 4' with the slope (dvu/dqz)r.
APP. A CONSTRUCTING CHARACTERISTIC CURVES 605
In computing the value of ( d y / d i , ~ in
~)~Eq. A.18a1 we use the approxi-
mate finite-difference form
Fig. A.2b. The values of cpz and (p, are given along the noncharac-
teristic curve 1-2, and either the magnitude or the direction of the veloc-
ity vector is given along the curves 1-4 and 2-3 (note that if curves 1-4
or 2-3 were solid boundaries, then the direction of the velocity would be
Collected in this Appendix are Tables B.l to B.8 for facilitating numerical
calculations.
Listed below are (i) an index to the tables, (ii) nomenclature for the tables,
(iii) the sources of the tables, and (iv) a bibliography of additional tables and
charts available in the literature.
INDEX TO TABLES
TABLE PAGE
B.1. Properties of the Standard Atmosphere . . . . . . . . . . . . . 612
B.2. Isentropic Flow, k = 1.4 . . . . . . . . . . . . . . . . . . . 614
B.3. Normal Shock, k = 1.4 . . . . . . . . . . . . . . . . . . . . 621
B.4. Frictional, Adiabatic, Constant-Area Flow (Fanno Line), k = 1.4 . 626
B.5. Frictionless, Constant-Area Flow with TO-Change (Rayleigh Line),
k=1.4.. . . . . . . . . . . . . . . . . . . . . . . . . 628
B.6. Influence Coefficients, k = 1.4 . . . . . . . . . . . . . . . . . 630
B.7. Hodograph Characteristic Functions for Two-Dimensional, Isen-
tropic, Supersonic Flow, k = 1.4 . . . . . . . . . . . . . . . 632
B.8. Coefficients for Second-Order Theory of Two-Dimensional, Supersonic
Flow, k = 1.4 . . . . . . . . . . . . . . . . . . . . . . . 633
SOURCES OF TABLES
B.1. T HE STAFF O F THE AMES 1- B Y %FOOTSUPERSONIC WIND T UNNEL SECTION.
Notes and Tables for Use in the Analysis of Supersonic Flow, NACA Tech.
Note, No. 1428 (1947).
B.2 to B.6. SHAPIRO, A. H., HAWTHORNE, W. R., and EDELMAN, G. M. The Mechan-
ics and Thermodynamics of Steady One-Dimensional Gas Flow with Tables
for Numerical Solutions, Meteor Report, No. 14, Massachusetts Institute
of Technology Guided Missiles Program, Cambridge, Mass. (1947).
B.7. STAFF OF THE G AS TURBINE LABORATORY, Massachuset.ts Institute of Tech-
nology, Cambridge, Mass. (1948).
B.8. Same source as for Table B.1.
610
APP. B BIBLIOGRAPHY OF TABLES AND CHARTS 61 I
3. KEENAN, J. H., and KAYE, J. Gr1.9 T a b h . New York: John Wiley & Sons, Inc..
1948.
4. SHAPIRO, A. H., HAWTHORNE, W. R., and EDELMAN, G. The Mechanics and
Thermodynamics of Steady One-Dimensional Gas Flow, in Handbook of
Supersonic Aerodynamics, NAVORD Report 1488 (Vol. I), Superintendent of
Documents, U. S. Govt. Printing Office (1950). Seealso EDELMAN, G . M., and
SHAPIRO, A. H. Tables for Numerical Solution of Problems in the Mechanics
and Thermodynamics of Steady, One-Dimensional Gas Flow Without D i s
continuities, Jour. App. Mech., Vol. 14, No. 4 (1947), p. A-344.
5. KOPAL, Z. Table of Supersonic Flow Around Cones, Tech. Rep., No. 1, M.I.T.
Center of Analysis, Mass. Inst. of Tech., Cambridge (1947).
6. KOPAL,2. Tables of Supersonic Flow Around Yawing Cones, Tech. Rep.,
No. 3, M.I.T. Center of Analysis, Mass. Inst. of Tech., Cambridge (1947).
7. KOPAL,Z. Supersonic Flow Around Cones of Large Yaw, M.I.T. Center of
Analysis, Mass. Inst. of Tech., Cambridge (1947).
8. ISENBERG, J. S. Monograph 11: The Method of Characteristics in Compressible
Flow; Part IA: Tables and Charts, Air Maten'el Command Report, No. F-TR-
1173-ND (Dec., 1947).
9. R O C K E J.~ ,A,, and HAYES, W. D. Monograph 11. The Method of Character-
istics in Compressible Flow; Part IC: Two-dimensional flow with large entropy
changes, Air Materiel Command Report, No. 102-AC49/6-100 (Dec., 1949).
10. NEICE, M. M. Tables and Charts of Flow Parameters Across Oblique Shocks,
NACA Tech. Note. No. 1673 (1948).
11. HUCKEL,V. Tables df ~ y ~ e r g e o h e t rFunctions
ic for Use in Compressible-Flow
Theory, NACA Tech. Note, No. 1716 (1948).
12. BURCHER. M. A. Compressible Flow Tables for Air, NACA Tech. Note, No.
1592 (1948).
13. T URNER , L. R., ADDIE,A. N., and ZIMMERMAN, R. H. Charts for the Analysis
of One-Dimensional Steady Compressible Flow, NACA Tech. Note, No. 1419
(1948).
14. AMES AERONAUTICAL LABORATORY. Notes and Tables for Use in the Analysis of
Supersonic Flow, NACA Tech. Note, No. 1428 (1947).
15. IVEY, H. R., STICKLE, G. W., and S C H U E ~ L EA. R , Charts for Determining the
Characteristics of SharpNose Airfoils in Two-Dimensional Flow a t Super-
sonic Speeds, NACA Tech. Note, No. 1143 (1947).
16. MOECKEL, W. E., and CONNORS, J. F. Charts for the Determination of Super-
sonic Air Flow Against Inclined Planes and Axially Symmetric Cones, NACA
Tech. Note, NO. 1373 (1947).
17. ORDNANCE AEROPHYSICS LABORATORY. Compressible Flow Tables and Graphs, in
Handbook of Supersonic Aerodynamics, NAVORD Report 1488 (Vol. 2),
Superintendent of Documents, U. S. Govt. Printing Office (1950).
18. ROSENHEAD, L. (ed.). Selection of TabEm for TJse in Calculations of Compressibk
Airflow. New York: Oxford Uiiversity Press, 1952.
TABLES OF COMPRESSIBLE-FLOW FUNCTIONS APP. B
TABLE B.l
PROPERTIES OF THE STANDARD ATMOSPHERE
-
Svmbols: See end of Table B.1.
App. B PROPERTIES OF THlE STANDARD ATMOSPHERE 613
TABLE B.2
ISENTROPIC FLOW
Perfect Gas, k = 1.4
-
Notes: (1) For values of M from 0 to 5.00, all digits to the left of the comma are
valid for linear interpolation. Where no comma is indicated in this
region, all digits are valid for linear interpolation.
(2) The notation .Ox429 signifies ,000429. The notation 53704 signifies
5,370.000.
APP. B ISENTROPIC FLOW 615
-
See Notes at beginning of this table.
618 TABLES OF COMPRESSIBLE-FLOW FUNCTIONS APP. B
---
See Notes at beginning of this table. ,
APP. B ISENTROPIC FLOW 619
TABLE B.3
NORMAL SHOCK
Perfect Gas, k = 1.4
Notes: (1) For values of M from 1.00 to 3.00, all digits to the left of the comma are
valid for linear interpolation. Where no comma is indicated in this
region, all digits are valid for linear interpolation.
622 TABLES OF COMPRESSIBLEFLOW FUNCTIONS APP. B
-~
TABLE B.4
FRICTIONAL, ADIABATIC, CONSTANT-AREA FLOW (FANNOLINE)
Perfect Gas, k = 1.4
~-
v/v*
h4 T/T* P/P* PO/PO* and F/F* 4.fLmdD
P*/P
APP. B FANNO LINE 627
FLOW (FANNO
TABLE B.4. FRICTIONAL, ADIABATIC, CONSTANFAREA L INE)
(Concluded)
Perfect Gas, k = 1.4
,
TIT* PIP* PO/PO* and FIF* 4fLrnanlD
P*/P
628 TABLES OF COMPRESSIBLE-FLOW FUNCTIONS APP. B
TABLE B.5
FRICTIONLESS, CONSTANT-AREA FWWWITH C HANGE IN STAGNATION TEMPERATURE
(RAYLEIQH LINE)
Perfect Gas, k = 1.4
P*/P
and
v/v*
APP- B RAYLEIGH LINE 629
TABLE B.6
-"
and an2
Fw/2 -Fw
M M~ 1 - M~ FA Ff and an'd
Fw/2 -Fw
5.0625
5.2900
5.5225
5.7600
6.0025
6.2500
6.5025
6.7600
7.0225
7.2900
7.5625
7.8400
8.1225
8.4100
8.7025
9.00
12.25
16.00
20.25
25.00
36.00
49.00
64.00
81.00
LOO. 00
m
632 TABLES OF COMPRESSIBLE-FLOW FUNCTIONS APP. B
TABLE B.7
w a a - o a a - w M
APP. B COEFFICIENTS FOR SECOND-ORDER THEORY 633
TABLE B.8
COEFFICIENTS FOR SECOND-ORDER THEORY OF TWO-DIMENSIONAL, SUPERSONIC
FLOW, k = 1.4
INDEX
(References are to pages of Vol. I only.)
Absolute temperature scale, 32 thickness drag of, 443
Accelerating flows, 507 trailing-edge flow of, 584
Acoustics, 48 two-dimensional, 567-80
Action. zone of. 500 types of, 575
~ d i a b a t i cellipse, 58-59 wake of, 578
Adiabatic flow wave drag of, 443
critical velocity for, 79-81 thin, supersonic flow for, 427
dimensionless velocitv M* of. 81-82 triangular profile, 574
energy equation of, 80-81 ' two-dimensional, 567-80
flow per unit area of, 82 Analogue techniques, 289
maximum velocity for, 79-81 Angular velocity, 269
one-dimensional, 160-1 78 Area change, flow with, 219 55
of perfect gas, 78-82 Area change and friction, flow with,
a t constant area, 162-73 24142
stagnation-temperature ratio for, 80 Area ratio, for isentropic flow, 85-86
Adiabatic Drocess. definition of. 29 Aspect ratio, effect on compressibility,
entropy chan g e' for, 34 400, 411-15
Adiabatic wall temperature, 211-12 generalized correlation of. 422
Affine transformation. 319 ~ r m o s ~ h e r&operti%
e, of, 612
Air, properties of, 612 Atomic explosion, 211
Airfoils; see also Wings Axially symmetric flow, 291,294-95,290,
biconvex, 568 596
completely symmetrical, 574
double-wedge, 568 Ballistics, 49
lift-drag polars of, 571 Barotropic fluid, 278
shock-characteristic theory of, 566 Bend. concave. 583
subsonic ~ e r n o u l lconstant,
i 276
pressure distributions for, 379-81 Bernoulli's equation, 276, 278
a t supersonic speeds, 571 Biconvex airfoil. 568
thickness of. 384-87 ~ i p l a n e supe&nic,
, 455, 579
in wind tunhel, 378 Body, free, 11
without stagnation points; see Bump Body, isolated, 11
supersonic Body forces, 18
aerodvnamicallv " ootimum.
* 575 Rodv of revolution: see Revolution.
biplane, 579 body of
center of pressure of, 444, 446, 574 Bomb, drag of, 408
downwash of. 576 Borda re-entrant orifice. 111
drag of, 442, ,446; 572 Boundary layer, 10, 26546, 282, 379
experimental data for, 4 4 7 4 8 displacement thickness of, 266
flow past, 501 and interaction with shocks, 135-37,
induced drag of, 442 266, 381-82, 580-84
lift of, 441, 445, 472 and pressure distribution, 308
lift-drag polar of, 446 separation of, 379, 380, 415
lift-drag ratio of, 443 Boundary of system, 11
linearized theory of, 4 4 0 4 9 Bump, flow past, 350, 372-74
of minimum thickness drag, 575 Burning, slow, 207-11
moment of, 444, 572
performance of, 445 Camber, in subsonic flow, 384-87
pressure drag of, 442 Cancellation of waves, 450, 452, 493
second-order theory of, 571-78 Cauchy-Riemann equations, 351
shape of, 571 Cavitation. 475
skin-friction drag of, 443 Center of ;ressure, of supersonic airfoils,
theory vs. experiment, 575 446, 572
INDEX
Centered waves, 474; see also Corner Compressibility burble, 326, 384
flow; Prandtl-Meyer wave effect. of aspect ratio on, 400
streamlines of, 525 Compressibility of liquid, 69
Chapman-Jouguet rule, 210-11 Compressible flow
Characteristic coordinates, 478 classification of, 58-59
Characteristic curves, 482-86, 598; see hypersonic, 59
also Method of characteristics subsonic, 59
eauations of. 602 supersonic. 59
hodograph, 599 transonic, 59
initial-value theorem of, 603-6 Compression shock; see Shocks
method of constructine. 603-6 Compression wave, steepening of, 127-3 1
orthogonality relationyor, 485, 603 Compressor, interference in, 315
physical, 599 Compressor cascades, 427, 462
properties of, 601 supersonic, 505
and shocks. 601 Concave bend, 583
as solut,ions, 601 Condensation
Characteristic function flow with, 219-55
series exoression of. 561 moisture. flow with: see S i r n ~ l eTo-
table of,'632 Change
Characteristic quadrilateral, 498, 499 Condensation shock, moisture, 203-5
Characteristics; see also Hodograph char- Conduct,ion of heat, Fourier law of, 36
acteristic: Method of characteristics Conductivitv. thermal. definition of. 36
hodograph, '467-68, 478, 485 ~onservatioh'of ma&; see also 'Con-
table of, 468 tinuity equation
working chart of, 467 principle of, 12
method of, 462-523; see also Char- for control volume, 13
acteristic curves; Method of char- Continuity equation, 14, 283-84; see also
acteristics Conservation of mass
illustrative example of, 479, 486-89 in cylindrical coordinates, 290
rules for using, 478 in spherical coordinates, 299
unit processes for, 492-94 in steady flow, 15
physical, 484 Continuum
physical significance of, 499 concept of, 4, 23
theory of, 482-91, 595-608 domain of, 56-58
Chemical reaction, flow with, 206-11, Euler's description of, 11
219-55; see also Simple To-Change Lagrange's description of, 11
Choking, 234-36, 257 limitations of, 56-58
due to heating, 192-93, 201-3, 205 mathematical description of, 10
due to Simple To-Change, 192-93, properties of, 5, 35
201-3, 205 Control surface, 12
in explosion wave, 210-11 Control volume, 12
frictional, 162, 170-78 First Law of thermodynamics for, 39
for isentropic flow, 89-94 Second Law of thermodynamics for, 40
for isothermal flow. 183 Converging nozzle
rules of, 235-36 discharge coefficient of, 93
thermal, 192-93, 201-3, 205 as flow meter, 93
Circular-arc profile, in subsonic flow, 375 as flow regulator, 93
Circular cvlinder with varying pressure ratio, 91-93
in subsonic flow, 367 Converging-diverging nozzle
in free jet, 368 operat,ing characteristics of, 1 3 9 4 3
in wind tunnel, 357-58, 367 with varying pressure ratio, 93-94
Circulation, 26749, 278-80, 284 Cooling, flow with; see Simple To-Change
Clausius, inequality of, 32 Corner flow, supersonic, 46344; see also
Coefficient Prandtl-Meyer wave; Simple-wave;
of discharge; see Discharge coefficient Centered waves
of friction; see Friction coefficient Critical Mach Number; see Lower critical
of heat transfer, 212-13 Mach Number
influence; see Influence coefficients Critical point,, 258
pressure; see Pressure coefficient Critical pressure ratio, 76
Combustion Crit,ical speed, 79-81
choking in, 198,201-2 Critical stsate,83
example of, 197-98 Critical velocity, in adishatic flow, 79-81
flow with, 206-11; see also Simple Crocco's thcorcm, 281-82
7'0-Change Crossing of waves; see Intersection of
Combustion chamber, 219 waves
INDEX 637
Equation of continuity; see Continuity Flow with waves of one family; see
equation Corner flow; Prandtl-Meyer flow;
Equations of motion, 265-97 Simple waves
Euler's, 273-78; see also Euler's equa- Fluid
tions barotropic, 278
linearized. 428 definition of, 3
Equipotenti'al lines homogeneous, 278
for flow past circular cylinder, 297 inviscid, 10
and streamlines. 295-97 Newtonian. 10
Euler. method of, '11 perfect, 10'
Euler's equations Fluid curve, 269
in Cartesian coordinates, 273-75 Fluid line. 269
integration of Fluid rot$tion. 269-73
for irrotational flow, 277-78 Flux
along a streamline, 275-77 of enthalpy, 39
in streamline coordinates, 281 of kinetic energy, 39
Evaporation, flow with, 219-55; see also of mass. 14
Simple To-Change of moment of momentum, 20
Exchanger, heat; see Heat exchanger of momentum, 17
Expansion wave, spreading of, 128-31 Force, units of, 20-21
Explosion Force break, 383-84
atomic, 211 Force-divergence Mach Number, 384-87
hvdrocarbon. 206-7 Forces
surf%e, 18
Fanno line, 114-15, 160-61, 192, 240 types of, 18
eouation of. 155 Fourier law of heat conduction, 36
tible of. 626 Friction
Field method, 491 one-dimensional flow with
Fields in su~ersonicflow. 477 adiabatic, 159-86, 2 1 S 5 5
Film coeffiiient of heat: transfer. 212- choking effects in, 170-78
curves for, 169
~ i r s t L a wof thermodynamics, 26 example of, 169-70
for control volume. 37-39 formulas for, 16649
for a cycle, 28 isothermal, 178-83; see &o Iso-
Flame front, 219 thermal flow in ducts
propagation speed of, 207, 209-11 laminar, 189
Flat late limiting duct length for, 170-71
noimal to subsonic stream, 360 in lone ducts. 173-78
subsonic flow past, 448 of perrect gas. 162-73
supersonic flow past, 444, 448 tables for, -169, 626
lift-drag ratio of, 446 simple, 159-78; see also Friction, one-
Fliegner's formula, 85 dimensional flow with
Flow; see also One-dimensional flow Friction and area change, flow with, 241-
accelerating, 507 42
adiabatic; see Adiabatic flow ~ r i c t & and heat transfer, flow with,
decelerating, 507 242-55
irrotational; see Irrotational motion Friction coefficient, definition of, 163
isentropic; see Isentropic flow experimental, 1 8 4 8 6
mass rate of, 14 for subsonic flow, 184-85
onedimensional for supersonic flow, 185-86
adiabatic, 160-78; see also Friction, Friction factor; see Friction coefficient
one-dimensional flow with, adi- Frictional effects on gas flow; see Fric-
abatic tion, one-dimensional flow with
with friction. 159-86; see also Fric- Frictional flow, table of, 626
tion, one-dimensional flow with Frictional heating, 211-12
potential; see Potential flow Frictionless flow, dimensional analysis of,
rotational; see Rotation; Rotational 308-9
flow
Flow coefficient, of nozzle, 99-100 Gas, perfect; see Perfect gas
Flow per unit area Gas constant, definition of, 41
isentropic, 84-85 and molecular weight, 41
maximum, 85 universal, 41
Flow with friction; see Friction, one- Gas constants, relations among, 42
dimensional flow witk Gas dynamics, 49
INDEX