You are on page 1of 49

November 26, 2008

MAE 3050

Review of Compressible Flow Topics


D. A. Caughey
Sibley School of Mechanical & Aerospace Engineering
Cornell University
Ithaca, New York 14853-7501
These notes provide a brief background to the thermodynamics of perfect gases and to compressible flow
theory. The notes are meant to complement the presentations in the text books used in M&AE 3050,
Introduction to Aeronautics, and M&AE 5060, Aerospace Propulsion Systems.

Chapter 1

Thermodynamic Review
1.1

Thermodynamic Equilibrium and Properties

There are many important concepts and results from the engineering science of thermodynamics that are
useful in fluid mechanics. Most of these results are based on the assumption that the substances with
which we deal are in thermodynamic equilibrium, which implies that all local mechanical, physical, and
thermal properties are independent of both position and time. Less is known about the thermodynamics
of nonequilibrium states, but numerous experiments have verified that the results for equilibrium states
approximate the behavior of the nonequilibrium states common in most fluid mechanical problems. Thus,
although the departures in the properties of moving fluids from equilibrium may seem large, their effect on
thermodynamic relations is, in fact, quite small.
Knowledge of thermodynamics is not required to understand the fluid mechanical behavior of liquids in
most cases. Important thermodynamic results relevant to gas flows are summarized here. A brief review of
thermodynamics is provided in Section 1.1.1.
Fluid properties such as pressure, specific volume, and temperature are thermodynamic properties, or state
variables. For a pure substance, any two state variables determine the thermodynamic state of the material,
so we can always, at least in principle, write an equation of state in the form
f (p, v, T ) = 0 ,

(1.1)

which relates the pressure p, specific volume v, and the temperature T . This equation of state can be written
equivalently as
p = fp (v, T ) .
(1.2)
The most common, and useful for aerodynamics, example of an equation of state is the perfect gas law, which
describes the behavior of gases at relatively low pressures and densities (relative to their critical points). The
perfect gas law takes the form
pv = RT ,
(1.3)
where R is the gas constant. In fluid mechanics, it is more common to work with the fluid density, which is
the inverse of the specific volume ( = 1/v), and so the perfect gas law can be written
p = RT .

(1.4)

See Section 1.2 for more details about the perfect gas law.
Pressure and density are defined more carefully in subsequent sections. The temperature is a state variable
that governs the tendency for heat to flow between two masses in thermal contact: heat always flows from
3

CHAPTER 1. THERMODYNAMIC REVIEW

the mass at higher temperature to the mass at lower temperature. If no heat is transferred when two masses
are brought into thermal contact, their temperatures must be equal.
The amount of heat that must be added to a substance to raise its temperature by a given amount is an
important property, called the specific heat. The specific heat can be determined either at constant volume,
in which case it can be identified as


e
,
(1.5)
cv =
T v
where e is the specific1 internal energy, and the subscript v indicates that the derivative is taken at constant
volume. The specific heat at constant pressure can be identified as


h
,
(1.6)
cp =
T p
where
h e + pv = e +

(1.7)

is the specific enthalpy, and the subscript p indicates that the derivative is taken at constant pressure.
The change in specific entropy s of a substance is defined as
s = s2 s1

dq
T

(1.8)

rev

where q is the heat added per unit mass during a reversible process between states 1 and 2. If the entropy is
introduced into an expression of the first law of thermodynamics, the following differential equation of state
results:
T ds = de + p dv = dh v dp .
(1.9)
Since only state variables appear in Eq. (1.9), this equation relates changes in these variables for a substance
whenever it is in thermodynamic equilibrium; i.e., it is valid even for processes that are neither reversible
nor adiabatic.

1.1.1

Review of Thermodynamics

The study of most problems in fluid mechanics involving liquids, and even those involving gases when the
pressure changes are not too large, does not require a knowledge of thermodynamics. This section is intended
to be a very brief review of the thermodynamic concepts required in the study of gas flows at high speeds,
which are treated in more detail in Chapter 2.
The thermodynamic state of any pure substance is determined uniquely by any two state variables. The
most common state variables used to describe fluids are the pressure p, the temperature T , and the specific
volume v. In fluid dynamics, it often is more convenient to work with the inverse of the specific volume,
which is called the fluid density:
1
= .
(1.10)
v
The first law of thermodynamics introduces the specific internal energy e. The first law states that work
and heat are distinguishable, but equivalent, forms of energy transfer and that the rate of change of internal
energy of a system consisting of a mass of fluid at rest is equal to the difference between the rates at which
1 The adjective specific is used to indicate the amount of a variable per unit mass. Thus, for example, while the internal
energy of a fluid is proportional to its mass, the specific internal energy, or internal energy per unit mass, is independent of the
total amount of mass in the system.

1.1. THERMODYNAMIC EQUILIBRIUM AND PROPERTIES

heat is added to, and work is done by, the system. If we represent these rates, per unit mass of fluid in the
system, as q and w, respectively, then the first law can be written
e = q w .

(1.11)

The most important way in which a fluid system does work on its surroundings is by the action of the fluid
pressure on the boundary of the system. If this work is done by a reversible process, then w = p dv, and
the first law can be written
e = q p dv .
(1.12)
The specific heat represents the amount of heat that must be added to raise the temperature of a unit mass
of the substance by a unit temperature
q
.
(1.13)
c=
T
Two particular versions of this quantity are of interest, depending on whether the process is carried out at
constant volume or at constant pressure. That is, we define the specific heat at constant volume
 
q
,
(1.14)
cv
T v
and the specific heat at constant pressure
cp

q
T

(1.15)

As seen directly from Eq. (1.12), the specific heat at constant volume is equivalent to


e
cv =
.
T v

(1.16)

Also, since from Eq. (1.12)


q = e + p dv ,

(1.17)

we see that if we define the specific enthalpy h as


h e + pv ,

(1.18)

q = dh v dp .

(1.19)

then Eq. (1.17) can be written in the form

From this it is seen that the specific heat at constant pressure is equivalent to


h
.
cp =
T p

(1.20)

The second law of thermodynamics implies the existence of another state variable, the specific entropy s,
defined such that in a reversible process the change in entropy is proportional to the heat added, with the
constant of proportionality being the inverse of the temperature. That is, for a reversible process the change
in specific entropy is given by
 
q
s
.
(1.21)
T rev
A direct consequence of this definition is that a reversible process that also is adiabatic takes place at constant
entropy; such a process is said to be isentropic.
Substitution of the definition Eq. (1.21) into the first law in the form of Eq. (1.17) gives
T ds = de + p dv .

(1.22)

CHAPTER 1. THERMODYNAMIC REVIEW

Since Eq. (1.22) contains only state variables, it must be independent of the path along which the process
takes place. Thus, it can be interpreted as a (differential) equation of state, describing how changes in
internal energy, entropy, and specific volume must be related. This equation of state also can be written in
terms of the density as
p
T ds = de 2 d .
(1.23)

Using Eq. (1.18) to replace the internal energy with the enthalpy also allows Eqs. (1.22) and (1.23) to be
written as
T ds = dh v dp ,
(1.24)
or
T ds = dh

dp
.

(1.25)

Equations (1.22), (1.23), (1.24), and (1.25) describe the behavior of any pure substance. In particular,
these equations can be used to relate changes in internal energy and density (or enthalpy and pressure) for
isentropic processes by setting ds = 0.
The implications of these equations for the behavior of gases that obey the perfect gas law are described in
Section 1.2.

1.2

Perfect gases

Over a fairly broad range of conditions, the equations of state for many gases are well approximated by the
equation of state for a perfect gas. The perfect gas law states that the pressure p, density , and temperature
T are related according to
p = RT ,
(1.26)
where R is called the gas constant. The gas constant for a particular gas can be determined from the
universal gas constant Ru by
Ru
R=
,
(1.27)
Mm
where Mm is the molecular weight of the gas. The value of the universal gas constant is
Ru = 8314.3

kg m2
.
s2 K mole

(1.28)

Thus, air, which has an average molecular weight of Mm = 28.964 kg/mole, has a gas constant of
Rair = 287.05

m2
.
s2 K

(1.29)

1.2. PERFECT GASES

1.2.1

Example: Perfect Gas Law Air

In the 1962 U. S. Standard Atmosphere [1], the temperature and pressure at 11,000 m above sea level are
T = 216.65 K and p = 22, 632 Pa, respectively. What is the corresponding density?
Solution:

Solving the perfect gas law for density gives


=

1.2.2

p
22, 632 Pa
=
= 0.36392 kg/m3 .
RT
(287.05 m2 /(s2 K))(216.65 K)

(1.30)

Example: Perfect Gas Law Uranium Hexafluoride

One of the heaviest gases of technical importance is uranium hexafluoride, for which the chemical formula
is U F6 . The molecular weight of this gas is 352 kg/mole. What are the values of the density and specific
gravity of this gas at standard sea-level conditions?
Solution:

The gas constant for uranium hexafluoride is


RUF6 =

Ru
8314.3 kg m2 /(s2 K mole)
=
= 23.62 m2 /(s2 K) .
Mm
352 kg/mole

(1.31)

Thus, the perfect gas law gives for the density


=

101, 325 Pa
p
=
= 14.89 kg/m3 .
RT
[23.62 m2 /(s2 K)](288.15 K)

The specific gravity is then


SG =

U F6
14.89 kg/m3
= 12.37 .
=
air
1.204 kg/m3

(1.32)

(1.33)

It is an interesting coincidence that the specific gravities of the heaviest known gas and the heaviest liquid,
mercury, (SGHg = 13.6) are approximately the same, when referenced to the most common gas (air) and
liquid (water), respectively.

1.2.3

Caloric Properties of Perfect Gases

The specific heat at constant volume cv for a perfect gas is a function only of the temperature T
cv = cv (T ) .

(1.34)

See Section 1.5 for proof that this follows directly from the equation of state for a perfect gas. Moreover, if
the gas is calorically perfect, i.e., if the constant-volume specific heat is constant, then the specific internal
energy e is given by
Z
e=

cv dT = cv T .

(1.35)

The corresponding specific enthalpy h, defined as

he+
can be written
h=

p
,

cp dT = cp T .

(1.36)

(1.37)

CHAPTER 1. THERMODYNAMIC REVIEW

The relationship between the specific heats at constant volume and at constant pressure for a calorically
perfect gas can then be seen from
h=e+

p
= cv T + RT = (cv + R)T = cp T .

(1.38)

Thus, for a calorically perfect gas, the difference in specific heats at constant pressure and constant volume
is given by
cp cv = R .
(1.39)
The Ratio of Specific Heats
The ratio of the specific heat at constant pressure to that at constant volume

cp
cv

(1.40)

plays a very important role in the behavior of compressible fluids. Introduction of this parameter into
Eq. (1.39) allows us to write
R
cv =
(1.41)
1
and

cp =

R
.
1

(1.42)

The ratio of specific heats is related to the number of modes in which energy can be stored by the molecules
of the gas. It is thus affected by both, the structure of the molecule and quantum effects. If the temperature
of the gas is such that m modes of energy storage are excited, then it can be shown that
m =

m+2
m

(1.43)

For a fairly broad range of common temperatures, the rotational energy modes of common molecules are
fully excited, but there is no significant energy in vibrational modes. In this case the value of m is the sum
of the three translational modes and the number of non-degenerate rotational modes:
m = 3 for monoatomic gases (such as helium), thus 3 = 5/3;
m = 5 for diatomic gases (such as oxygen and nitrogen), thus 5 = 7/5;
m = 6 for (non-collinear) triatomic gases (such as carbon-dioxide and water), thus 6 = 8/6.
For further information about the values of for various gases, see any common thermodynamics textbook
(e.g., [3]).

1.3

Compressibility

All fluids experience changes in density with changes in pressure. For liquids the changes usually are quite
small, but for gases the changes can be very large. These changes in density are equivalent to changes in
specific volume, since the density is the inverse of the specific volume v:

1
.
v

(1.44)

1.3. COMPRESSIBILITY

The compressibility of a substance represents the fractional amount by which its volume changes in response
to changes in pressure. Thus, the compressibility is defined as
lim

p0

v/v
1 dv
=
,
p
v dp

(1.45)

where v = 1/ is the specific volume. The minus sign is introduced to make positive, since an increase in
pressure generally results in a decrease in specific volume. By virtue of Eq (1.44), an equivalent definition
of the compressibility is
1 d
.
(1.46)

dp
Note that the minus sign is absent from Eq. (1.46), since a decrease in specific volume corresponds to an
increase in density.
The definitions in Eqs. (1.45) and (1.46) are not complete until we specify how the temperature (or some
other state variable) changes during the compression process. Two specifications are commonly used. The
isothermal compressibility T is defined for a process in which the temperature is held fixed
 
1 v
T
,
(1.47)
v p T
while the isentropic compressibility s is defined for a process in which the entropy is held fixed
 
1 v
s
.
v p s

(1.48)

A thermodynamic analysis shows that the ratio of the isothermal compressibility to the isentropic compressibility is equal to the ratio of the specific heat at constant pressure cp to the specific heat at constant volume
cv :
cp
T
=
=.
(1.49)
s
cv
For liquids, there is only a very small change in temperature with pressure in isentropic processes, and the
isothermal value of the compressibility usually is used. The isothermal compressibility of liquids is a weak
function of temperature that usually must be determined empirically. It also is common to work with the
reciprocal of the compressibility, which is called the bulk modulus of elasticity E, defined as
 
 
1
p
p
=
=
.
(1.50)
E v
v T
T
T
From Eq. (1.50), it is obvious that the dimensions of the bulk modulus are the same as those of pressure.
The dependence of liquid densities on pressure for an isentropic process is reasonably well approximated by
the empirical relationship
1/n

p/pref + B

,
(1.51)
=
ref
B+1
where B and n are constants. The values B = 3, 000 and n = 7 provide a good fit for both fresh and sea
water when atmospheric reference conditions are chosen [2]. Differentiation of Eq. (1.51) gives the isentropic
compressibility at the reference conditions p = pref and = ref as
s =

1
.
n(B + 1)pref

(1.52)

The bulk modulus of water at one atmosphere pressure is a weak function of temperature and passes through
a maximum value of E = 2.292 109 Pa at a temperature T = 49o C. The bulk modulus of water at one
atmosphere pressure and T = 20o C is E = 2.182 109 Pa.

10

1.3.1

CHAPTER 1. THERMODYNAMIC REVIEW

Example: Compressibility of Water

To what value must the pressure be increased from one atmosphere to increase the density of water at
T = 20o C by 1%?
Solution:

Equation (1.51) can be solved for the pressure ratio to give



n
p

= (B + 1)
B.
pref
ref

(1.53)

Thus, a 1% increase in density requires


p
n
= (B + 1) (1.01) B = 3001(1.01)7 3000 = 217.5
pref

(1.54)

p = 217.5 pref = 217.5 atm .

(1.55)

whence
Alternatively, since the density change is so small, we can approximate Eq. (1.46) as
/
1

= s ,
p
p

(1.56)

which can be solved for the pressure change p


p =

1
,
s

(1.57)

using the value of isentropic compressibility given by 1/s = n(B + 1)pref , corresponding to the reference
conditions. Thus, for a 1% change in density
p =

0.01
= 0.01n(B + 1)pref = 0.01(7)(3, 001)pref = 210 atm .
s

(1.58)

This approximation gives a slightly smaller value for the required pressure increase because it does not
account for the increase in compressibility with pressure. Either result indicates that the pressure must be
increased to more than 200 atm to produce a 1% change in density of water. Thus, the compressibility of
water must be taken into account only for extremely large pressure changes; for most flows, water can be
treated as a fluid whose density is independent of the pressure.
As is seen in Example 1.3.1, changes in the density of water due to changes in pressure are very small
unless the pressure changes are very large. Thus, it is usually a good approximation to treat water as an
incompressible fluid. It is important to emphasize, however, that does not necessarily mean that the density
is constant. Even though the compressibility of water usually can be neglected, changes in the density of
water can become important in certain stratified flows. Such density changes usually are the result of changes
in temperature and/or the amount of dissolved solids. This results in incompressible flows in which it is
necessary to account for changes in fluid density.
For gases, the compressibility can be computed from the equation of state. For perfect gases, which obey
the equation of state
p = RT ,
(1.59)
the isothermal compressibility is
T =

1
.
p

(1.60)

Equation (1.49) then shows that the isentropic compressibility for a perfect gas is
s =

1
.
p

(1.61)

11

1.3. COMPRESSIBILITY

The compressibility is related to the speed of sound in the fluid. In Chapter 2 it is shown that the speed of
sound a is given by
s 
r
1
p
=
.
(1.62)
a=
s
s
Since the isothermal and isentropic compressibilities of liquids are nearly the same, we can write
r
p
1
E/
a=
s

(1.63)

for liquids, while for a perfect gas,

a=

RT =

p
.

(1.64)

Thus, the speed of sound in a perfect gas is proportional to the square root of the temperature.

1.3.2

Isentropic Processes involving Perfect Gases

Flows in which the effects of heat transfer and of fluid friction are small are well approximated as isentropic
processes. The differential equation of state for a pure substance relates the changes in entropy s to those
in enthalpy h and those in density according to
T ds = dh

dp

(1.65)

Thus, for a constant entropy (or isentropic) process ( ds = 0) involving a perfect gas, we can write
dp
= cp dT



R
p

d
1
R

dh

dp
= 0,

dp
= 0.

(1.66)

Since R is constant, this can be simplified to

which can be integrated to give

d
dp

=0
p

(1.67)

p
= Const.

(1.68)

This important equation relates the pressure and density for any isentropic process involving a perfect gas.
For any such isentropic process, Eq. (1.68) and the equation of state Eq. (1.26) can be used to relate the
changes in any one state variable to those of only one other state variable:
 
p

=
(1.69)
p1
1
 /(1)
T
p
=
(1.70)
p1
T1
 1/(1)

T
=
(1.71)
1
T1

12

1.3.3

CHAPTER 1. THERMODYNAMIC REVIEW

Example: Isentropic Processes

What percentage pressure change is required to produce an increase in density of 5 per cent for an isentropic
process in air?
Solution:

Assuming the process is isentropic, and that for air = 1.4, we have


p + p
+
1.4
= (1.05) = 1.0707
=
p

(1.72)

Thus, an increase of approximately 7 per cent is required. Under standard sea level conditions, this corresponds to a pressure change of approximately 7,165 Pa (or 1.04 psi). Note that since the pressure and density
changes are small, these could have been predicted approximately by Eq. (1.67), which can be written as

= 1.4(0.05) = 0.07
p

1.4

(1.73)

The Speed of Sound

The speed at which infinitesimal pressure disturbances propagate in a fluid, i.e., the speed of sound, plays
a very important role in the flow of compressible fluids. The conservation laws of mass and momentum are
used in Chapter 2 to show that the speed of sound is equal to
a2 =

dp
.
d

(1.74)

In addition, the processes occurring sound waves of all but the highest frequencies are slow enough to be
nearly reversible, and the thermal conductivities of common gases, such as air for a broad range of densities,
are sufficiently small that the process of sound propagation also is very nearly adiabatic, so that sound waves
are very nearly isentropic. Thus we can write the equation determining the sound speed as
 
p
2
,
(1.75)
a =
s
where the subscript ()s denotes the isentropic derivative of pressure with respect to density. Finally, for a
calorically perfect gas undergoing an isentropic process it is seen from Eq. (1.67) that
d
dp

= 0,
p

so we can write
2

a =

=
s

p
= RT ,

(1.76)

(1.77)

where the final form was obtained using the equation of state for a perfect gas, Eq (1.26). Thus, the sound
speed for a calorically perfect gas is proportional to the square root of its temperature T .
That the speed of sound should be proportional to the square root of the temperature of a gas is consistent
with a simple kinetic theory picture of the gas at the molecular level. This is because the the temperature
is a measure of the kinetic energy of the random molecular motion of the molecules, whence the mean speed
of this motion is proportional to the square root of the temperature, and it is reasonable to assume that the
speed at which disturbances are propagated through the gas should be proportional to this mean speed.

1.5. APPENDIX: SPECIFIC HEATS OF CALORICALLY PERFECT GASES

1.5

13

Appendix: Specific Heats of Calorically Perfect Gases

The Helmholtz function is defined as


auT s

(1.78)

da = p dv s dT

(1.79)

whence, using Eq. (1.22), we can write


This equation is equivalent to stating that

and

a
v

a
T

= p

(1.80)

= s

(1.81)

Since, for sufficiently differentiable functions the mixed cross derivative is independent of the order in which
the derivatives are taken, these two equations lead to


s
p
=
(1.82)
T v
v T
This one of the four Maxwell relations; the other three come from equivalent operations on the differential
equations of state, Eqs. (1.22) and (1.24), and the corresponding formula for the differential of the Gibbs
function.
Now, if we assume that the internal energy is a function of specific volume v and temperature T , we have


u
u
du =
dv +
dT
(1.83)
v T
T v
and, the differential equation of state, Eq. (1.22), tells us that


u
s
=T
v T
v T
Substituting from Eq. (1.82) gives


u
v

p
=T
T

and, since the equation of state for a perfect gas gives



p
p
= R =
T v
T
we have

u
v

=T

p
p=0
T

Thus, the internal energy of a perfect gas must be a function only of its temperature.

(1.84)

14

CHAPTER 1. THERMODYNAMIC REVIEW

Bibliography
[1] Anon., [1962]. U. S. Standard Atmosphere, 1962, U. S. Government Printing Office, Washington.
[2] R. H. Cole, [1948]. Underwater Explosions, Princeton University Press.
[3] John R. Howell & Richard O. Buckius, Fundamentals of Engineering Thermodynamics, Second
Edition, McGraw-Hill, 1992.

15

16

BIBLIOGRAPHY

Chapter 2

Summary of Results from


Compressible Flow Theory
This chapter provides a summary of some useful results from compressible flow theory, concentrating on
isentropic flows and flows with shock waves. Further details can be found in the classic text books [4] and
[6].

2.1

Steady, Adiabatic Flows

The energy equation for steady, inviscid, adiabatic flow can be derived either from the Bernoulli equation
or from the First Law of thermodynamics. When it is derived from the Bernoulli equation, we must include
the assumption that the flow is isentropic while, when it is derived from the First Law, we require only the
assumption that the flow be adiabatic i.e., that there is no heat transfer to the fluid.

2.1.1

Compressible Bernoulli Equation

We first consider the compressible form of the Bernoulli equation. The stream wise component of the
Momentum Theorem, for the control volume illustrated in Fig. 2.1, gives the differential form of the Bernoulli
equation
dp + V dV = 0 ,

(2.1)

for steady flow along a streamline. In this equation, p and are the fluid pressure and density, respectively,
and V is the velocity; the gravitational term has been neglected, as is usual in aerodynamics.
If the flow is compressible, changes in density must be taken into account when integrating this equation.
If we assume that changes in density are related to those in pressure by the equation describing an isentropic
process i.e., Eq. (1.69) then the integral of this equation becomes
p V2
+
= Const. ,
1
2

(2.2)

where = cp /cv is the ratio of specific heats. This is the compressible form of the Bernoulli equation,
relating pressure (and density!) to velocity.
17

18

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY


p + dp
e + de
V + dV

p, e, V, A

A + dA

+ d

ds

Figure 2.1: Differential element of a stream tube for deriving the compressible Bernoulli and energy equations.
The perfect gas law can be used to replace the ratio of pressure to density in the compressible Bernoulli
equation with the product RT of the gas constant and temperature and, since
R
= cp ,
1
Eq. (2.2) can be written
V2
= Const.
2
Finally, for a calorically perfect gas, for which cp T = h, the enthalpy, we have
cp T +

h+

V2
= hT ,
2

(2.3)

where hT is the total, or stagnation, enthalpy. As will be seen in the next section, this is the energy equation
for a steady, adiabatic, but not necessarily reversible, flow.

2.1.2

Steady Energy Equation

The First Law of thermodynamics for steady flow through a control volume can be written

ZZ 
V2 ~
V n
dS ,
QW =
e+
2

(2.4)

where e is the specific internal energy, Q is the rate at which heat is being added to the fluid in the control
volume, and W is the rate at which the fluid in the control volume is doing work on the surroundings.1
Consistent with the treatment of the preceding section, we neglect the work done by the gravitational force.
For the control volume consisting of the differential element of stream tube illustrated in Fig. 2.1,
ZZ
~ n
W =
pV

and we will assume there is no heat transfer, i.e., that Q = 0. For the given control volume, Eq. (2.4) then
becomes


1
1 2
2
pAV (p + dp)(A + dA)(V + dV ) = AV e + de + (V + dV ) (e + V ) .
(2.5)
2
2
The Continuity Equation applied to the same control volume gives


d
AV
= AV 1
+ ,
(A + dA)(V + dV ) =
+ d

1 Here,

unlike Section 1.1.1, the total energy is taken to include the kinetic energy of the flowing gas.

(2.6)

19

2.1. STEADY, ADIABATIC FLOWS

V=a

V = a + dV

p + dp

+ d

T + dT

(a)

(b)

Figure 2.2: Determination of the speed of sound. (a) Planar sound wave traveling at speed a through
stationary atmosphere; (b) Sound wave and control volume enclosing it in reference frame traveling with
wave.
which can be used to simplify Eq. (2.5) to




V2
p V2
= d h+
= 0,
d e+ +

2
2

(2.7)

or

V2
= hT ,
(2.8)
2
where the stagnation enthalpy hT is constant for a given flow. This equation is seen to be identical to the
energy form of the compressible Bernoulli equation, Eq. (2.3). Our assumption that the flow is inviscid means
that work done by frictional forces acting on the sides of the stream tube is negligible, but no assumptions
have been made about whether there are irreversible processes inside the control volume. In particular,
this equation will hold if the control volume contains shock waves which, as will be seen in Section 2.3, are
irreversible (but adiabatic).
h+

2.1.3

Sound Speed in a Continuous Medium

We consider the one-dimensional process corresponding to an infinitesimal pressure wave propagating through
a fluid that is initially at rest. The situation is as shown in Fig. 2.2(a). This flow pattern is unsteady, since
the pressure wave is moving with velocity a relative to our reference frame (which is attached to the fluid at
rest). The problem can be viewed as a steady one, provided the problem is transformed to a reference frame
moving with the pressure wave; in this case the flow becomes that shown in Fig. 2.2(b).
We apply the laws of conservation of mass and momentum to the control volume shown by the broken lines
in Fig. 2.2(b). The sides of the control volume are aligned with the flow, so the only fluxes are across the
faces having area A. Conservation of mass thus requires
aA = ( + d)(a + dV )A

(2.9)

0 = dV + a d

(2.10)

or
where the higher-order differential d dV has been neglected, since it will vanish in the limit of infinitesimal
disturbances.
The momentum theorem, applied to the same control volume, gives for the x-component
(p + dp)A pA = Aa2 + ( + d)A(a + dV )2 .

(2.11)

20

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

Using the continuity equation, this expression can be simplified to


(p + dp)A pA = Aa dV ,

(2.12)

dp + a dV = 0 .

(2.13)

or
Note, since a = V , the fluid velocity entering the control volume, this is simply equivalent to the Bernoulli
equation, in differential form, for inviscid flow along a streamline.
Now, if the results from the continuity and momentum equations both are solved for dV (to eliminate this
quantity from the final expression) we find
dV =

a d
dp
=
a

which, in turn, can be solved for a2 to give


a2 =

dp
.
d

(2.14)

(2.15)

Our mechanical analysis here does not specify what is held constant in taking this derivative, but the
viscosities and thermal conductivities of common gases, including air, are sufficiently small that the process
is nearly adiabatic and reversible, hence isentropic. Thus, as seen earlier in Chapter 1,

p
dp
2
=
= RT .
(2.16)
a =
d s

2.1.4

Example: Speed of Sound

Determine the sound speed for air at standard sea level conditions. What is the speed of sound in hydrogen
at the same temperature?
Solution: The temperature of air under standard sea level conditions [1] is T = 288.15 K, so the speed of
sound under these conditions is
p
p
a = RT = (1.4)(287.05)(288.15) m2 /s2 = 340.3 m/s .
(2.17)
The speed of sound of gases depends on their molecular weights. The molecular weight of hydrogen is
MH2 = 2, so its gas constant is
RH2 = Ru /MH2 = (8314/2) m2 /(K s2 ) = 4157 m2 /(K s2 ) .

(2.18)

Thus, the speed of sound in hydrogen at the temperature corresponding to standard sea level conditions is
q
p
2
(2.19)
a = RH2 T = (1.4)(4157)(288.15) m2 /s = 1295 m/s .

2.1.5

Stagnation Temperature, Pressure, and Density

For a calorically perfect gas, the enthalpy can be expressed as h = cp T , and Eq. (2.8) can be rearranged to
give
1 2
TT
=1+
M ,
(2.20)
T
2

21

2.1. STEADY, ADIABATIC FLOWS

where TT is the stagnation temperature of the flow and T is the static temperature at a point in the flow
where the local Mach number is M = V /a, where a is the speed of sound. In developing Eq. (2.20) we have
used the fact that the square of the speed of sound is given by a2 = RT (see Sections 1.4 and 2.1.3).
The isentropic stagnation pressure pT and the isentropic stagnation density T are defined as the values of
pressure and density, respectively, achieved when the flow is brought to rest (i.e., to stagnation conditions)
by a reversible and adiabatic process. For such an isentropic process, we can combine the equations relating
pressure and density to temperature for isentropic processes with the energy equation in the form of Eq. (2.20)
to give
/(1)
 /(1) 
1 2
TT
pT
= 1+
,
(2.21)
=
M
p
T
2
and

T
=

TT
T

1/(1)

1+

1 2
M
2

1/(1)

(2.22)

Note that the stagnation temperature is achieved when the flow is stagnated by any adiabatic process, while
the stagnation pressure and density are achieved only when the stagnation process also is reversible. Any
irreversibility in the process will cause a loss in stagnation pressure (and stagnation density).

2.1.6

Example: When is a Flow Incompressible?

We might reasonably expect that a flow can be approximated as incompressible so long as the ratio of
stagnation density to density in the free stream is less than 1.10; to what maximum Mach number does this
correspond?
Solution:

We can solve Eq. (2.22) for the Mach number to give


M=

2
1

"

1

#)1/2

Thus, for the specified ratio of stagnation to static density, and assuming = 1.4 for air,

i1/2
2 h
0.4
(1.10) 1
Mmax =
= 0.44
0.4
Airspeed measurements sometimes interpret the pressure difference pT p as the dynamic pressure V 2 /2,
using the incompressible Bernoulli equation. The ratio
"
#
/(1)
2
1 2
pT p
=
1+
1
(2.23)
M
1
2
M2
2
2 V
is plotted as a function of free stream Mach number in Fig. 2.3. As expected, the ratio approaches unity in
the limit as M 0. Furthermore, it is within 5 per cent of unity for Mach numbers less than M = 0.44. That
is, using the incompressible Bernoulli equation to infer the dynamic pressure from the Pitot-static pressure
difference incurs less than 5 per cent error at Mach numbers up to M = 0.44. The binomial theorem can be
used to expand the first term in the square brackets of Eq. (2.23), giving the asymptotic expansion
M2
2 4
pT p
=
1
+
+
M + O(M6 ) .
1
2
4
24
V
2

(2.24)

This result also is plotted in Fig. (2.3), and is seen to closely approximate the exact result for Mach numbers
as large as M = 2.0.

22

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

1.35

6
Exact
Asymptotic

Exact
Asymptotic

5.5

1.3
Dynamic pressure ratio, (p p)/q

4.5

Dynamic pressure ratio, (p p)/q

5
1.25

1.2

1.15

1.1

4
3.5
3
2.5
2

1.05
1.5
1

0.1

0.2

0.3

0.4
0.5
0.6
Mach number, M

0.7

0.8

0.9

0.5

1.5
Mach number, M

2.5

Figure 2.3: Ratio of the pressure difference pT p to the dynamic pressure 12 V 2 as a function of free stream
Mach number.
Often it is convenient to use sonic conditions as a reference, rather than stagnation values. We use the
superscript asterisk to denote sonic conditions, so V = a , by definition, where a is the speed of sound.
Equations (2.20), (2.21), and (2.22) show that the ratio of any thermodynamic variable at the sonic point to
its stagnation value is a function only of the ratio of specific heats . That is, setting M = 1 in Eqs. (2.20),
(2.21), and (2.22) yields
TT
+1
=
,
(2.25)
T
2
/(1)

+1
pT
,
(2.26)
=
p
2
and
T
=

2.2

+1
2

1/(1)

(2.27)

Isentropic Flow in Ducts of Varying Area

The properties of steady, isentropic flows can be expressed as functions of the Mach number and the ratio
of the stream tube (or duct) cross-sectional area to the area, real or virtual, at which the flow is sonic. For
steady flow, the continuity equation can be written as
AV = A V ,
where the asterisk
re-written as

(2.28)

again refers to sonic conditions. Since, by definition, V = a this equation can be

or

a 1
A

= ,
a M
A

(2.29)

A
T a aT 1

= .
T aT a M
A

(2.30)

Finally, recalling that aT /a = (TT /T )1/2 and using Eqs (2.20), (2.21), (2.22) and Eqs. (2.25), (2.26),
and (2.27), this relationship can be written as
1
M

2
1 2
+
M
+1 +1

+1
 2(1)

A
.
A

(2.31)

23

2.2. ISENTROPIC FLOW IN DUCTS OF VARYING AREA

1
T/T

p/p

0.9

0.8

A /A
0.7

Ratio

0.6
0.5
0.4
0.3
0.2
0.1
0

0.5

1.5

2
2.5
3
Mach number, M

3.5

4.5

Figure 2.4: Area ratio and flow variables as functions of Mach number for isentropic flow; calorically perfect
air ( = 1.4).
This relationship and the isentropic ratios of the flow variables to their stagnation values given by Eqs (2.20),
(2.22), and (2.22) completely determine the properties of isentropic flows in ducts (or stream tubes) of slowly
varying cross-sectional area; these results are presented in Fig. 2.4.
A consequence of these equations is that the mass flow rate m
per unit area for given stagnation conditions
is a function only of the Mach number M (and, parametrically, the ratio of specific heats ). That is,
m

V a
= V =
T aT
A
T a aT
#
"r

pT
M
=
R 1 + 1 M2 (+1)/(2[1])
TT
2
#
"

M
pT
=
(+1)/(2[1])
RTT
1 + 1 M2

(2.32)

The quantity in square brackets in the next to last line of the above equation is called the mass flow parameter
MFP by Mattingly [5]
r

M
MFP =
(2.33)
R 1 + 1 M2 (+1)/(2[1])
2

I prefer to define the related quantity MFP, which is independent of the gas constant R

M
MFP =
(+1)/(2[1])
1
2
1+ 2 M

(2.34)

Thus, we can express the mass flow rate as

pT A
pT A
MFP(M, )
m
= MFP(M, R, ) =
TT
RTT

(2.35)

The mass flow parameter MFP and the ratio A/A are inverse measures of the same quantity. In fact,
comparison of Eqs. (2.31) and (2.34) shows that
(+1)/(2[1])

1
2
A

=
(2.36)

A
+1
MFP

24

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

2.5

VA
VA
A

a/a

V/a
*
/
Mach

2.5

2
1.5

1.5
1

0.5

0.5

0
0

0.5

1.5

2.5

0.5

1.5

V/a*

(a)

2
Mach

2.5

3.5

(b)

Figure 2.5: Flow variables as functions of flow velocity or Mach number for isentropic flow; calorically perfect
air ( = 1.4). (a) Speed of sound and other flow variables as functions of flow velocity; (b) Continuity equation
and its low-Mach and high-Mach approximations; all quantities normalized by their sonic values.
The isentropic flow properties are given in tabular form as functions of Mach number for both, subsonic
and supersonic flows, in Tables 2.1 and 2.2 in Section 2.8 for a calorically perfect gas having ratio of specific
heats = 1.4 (representative of air).

2.2.1

The Adiabatic Ellipse

A useful summary of how flow properties change with Mach number, illustrating the qualitative differences
between subsonic and supersonic flow, can be developed from the energy equation (see Eq. (2.8)). Since the
enthalpy of a calorically perfect gas can be expressed as
h = cp T =

a2
RT
=
1
1

where, as usual, a represents the speed of sound, Eq. (2.8) can be written as
 a 2 1  V 2
+1
,
=
+
a
2
a
2

(2.37)

where a again represents the speed of sound at the sonic point. This result is called the adiabatic ellipse,
since it shows that a plot of the speed of sound a as a function of the fluid velocity V is an ellipse.
The adiabatic ellipse is shown in Fig. 2.5(a) for calorically perfect air ( = 1.4). Note that a steady, adiabatic
flow has a maximum achievable velocity, given by
r
 
V
+1
=
.
(2.38)
a max
1
This corresponds to the limit when T 0 (and, correspondingly a 0), i.e., when all the thermal energy
has been converted to directed, kinetic energy. Note that, by virtue of the shape of the ellipse, the speed

25

2.3. NORMAL SHOCK WAVES

of sound approaches a constant as the velocity goes to zero, while the velocity approaches a constant as the
speed of sound goes to zero.
The quantity
M =

V
a

(2.39)

is simply a normalized velocity, although it has some properties in common with the Mach number M = V /a.
In particular, M < 1 when the Mach number is subsonic, and M > 1 when the Mach number is supersonic.
Note, however, that M has a finite maximum, corresponding to the maximum achievable velocity. The Mach
number, however, approaches infinity in this limit, since the speed of sound goes to zero there. The Mach
number is also plotted as a function of V /A in Fig. 2.5(a). Note that for subsonic Mach numbers, M is
not very different from the Mach number M since the speed of sound changes by less than 10% between
stagnation conditions and the sonic point but, of course, it deviates significantly for supersonic flows.
Finally, the variable / is also plotted in Fig. 2.5(a). As expected, the density approaches a constant,
consistent with the incompressible approximation, in the limit as the velocity goes to zero. Recall that the
sonic point is the critical point where changes in density become more important than changes in velocity in
the continuity equation (as the flow transitions from subsonic to supersonic Mach numbers). This is reflected
in the fact that the slope of the density curve is equal, but of opposite sign to the slope of the velocity curve
at the sonic point in Fig. 2.5.
The various factors involved in the steady continuity equation are plotted in Fig. 2.5(b); for all three curves,
the variables are normalized by their sonic values. The product V A is seen to be constant for all Mach
numbers, as required by the continuity equation. In the limit of small Mach number, the product V A is seen
to approach a constant, since the density is nearly constant, at its value

T / =

+1
2

1/(1)

And, in the limit of large Mach number, the product A is seen to approach a constant, since the velocity
V /a is nearly constant, at its value
Vmax /a =

( + 1)/( 1) .

These two observations confirm that, in order to satisfy the steady-flow continuity equation (V A = const.),
at small Mach numbers changes in cross-sectional area are balanced primarily by changes in velocity (i.e.,
V A const.); while at large Mach numbers changes in cross-sectional area are balanced primarily by
changes in density (i.e., A const.).

2.3

Normal Shock Waves

The governing laws of conservation of mass, momentum and energy, applied to a control volume containing
a shock, require
1 V1 = 2 V2 = m
,
(2.40)
p1 + 1 V12 = p2 + 2 V22 ,
h1 +

V12
V2
= h2 + 2 = hT .
2
2

(2.41)
(2.42)

Note that the stagnation enthalpy is constant, since we are assuming the process is adiabatic i.e., that no
heat is added but the flow will not necessarily be isentropic, since we have not assumed the flow to be
reversible.

26

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

30

p2/p1

p2/p1

/
2

25

T /T
2

T /T
2

1.5

2.5
3
3.5
Mach number, M

4.5

10

15

p /p

Ratio

Ratio

p /p

20

1.5

2.5
3
3.5
Mach number, M

4.5

Figure 2.6: Shock relations for calorically perfect air (with = 1.4). The same data are plotted in both
figures; the vertical scale of the figure on the right is chosen to clarify the behavior for small M1 .
The ratios of the thermodynamic variables on either side of a normal shock depend only upon the normal
Mach number M1 of the flow upstream of the shock (and the ratio of specific heats ) for the case of a
calorically perfect gas. These can be written in the form

p2
2
=1+
M21 1 ,
p1
+1
2
=
1
1+
M22

T2
=
T1

n
1+

2
+1

1+
1+

(2.43)

M21
,
(M21 1)

(2.44)

1
+1

1
+1
2
+1


M21 1

(M21 1)

o n
1+
M21 1
M21

1
+1

(2.45)
o
M21 1

(2.46)

Equation (2.45) gives values of M2 < 1 when M1 > 1, and values of M2 > 1 when M < 1; i.e., a normal shock
wave must convert a supersonic flow to a subsonic one, and vice versa. Equation (2.43) can be combined
with Eq. (2.21) to give an expression for the ratio of stagnation pressures across the shock, namely
pT2
=
pT1

( + 1)M21
2 + ( 1)M21

/(1) 

+1
2M21 ( 1)

1/(1)

(2.47)

The Second Law of Thermodynamics requires that the stagnation pressure not increase across a shock, and
inspection of Eq. (2.47) shows that this requires that M1 not be less than unity. Thus, the Second Law limits
possible shock waves to only those having M1 1. For these shock waves, as we have seen, we will have
M2 1 thus, only shock waves across which the flow changes from supersonic to subsonic velocities are
allowed. For these shock waves, the pressure, density and temperature all must increase across the shock.
The relationships for the behavior of normal shocks in Eqs. (2.43) (2.47) are plotted in Fig. 2.6 for calorically
perfect air (with = 1.4); consistent with the discussion in the preceding paragraph, the curves are plotted
only for supersonic upstream Mach numbers M1 1.

27

2.3. NORMAL SHOCK WAVES

30
p /p
t

25

pt /p1
1

Ratio

20

15

10

0.5

1.5

2
2.5
3
Mach number, M

3.5

4.5

Figure 2.7: Rayleigh Pitot formula compared with isentropic compression as a function of flight Mach number
M1 ; calorically perfect air with = 1.4.
A properly designed Pitot-static probe, when immersed in a supersonic stream, measures the ratio of stagnation pressure downstream of a normal shock to the upstream static pressure; this is given by
1/(1)

/(1) 
pT2
+1 2
+1
.
(2.48)
=
M1
p1
2
2M21 ( 1)
This equation is known as the Rayleigh Supersonic Pitot formula, and is the basis for calibrating Mach
meters for supersonic flight. Figure 2.7 compares the ratio pT2 /p1 with the isentropic ratio pT1 /p1 ; it is seen
that there is very little loss in stagnation pressure for flight Mach numbers less than about M1 = 1.3, but
the differences become very significant at larger Mach numbers.
In addition to these general results, several other features of shock waves are evident from Fig. 2.6. Note,
for weak shock waves, the stagnation pressure ratio is very nearly equal to unity. In fact, it can be shown
from Eq. (2.47) that

3
4

( + 1)2 p
p
pT2
=1
+
O
,
(2.49)
pT1
12 2
p1
p1

where p = p2 p1 is the pressure change across the shock. That is, the loss in stagnation pressure is
of third order in the strength of the shock. This means that for small values of the pressure change (i.e.,
p/p1 << 1) the change in stagnation pressure will be extremely small; i.e., weak shocks are very nearly
isentropic.

It also is evident from Fig. 2.6 that both the density ratio 2 /1 and the downstream Mach number M2
approach finite values as the upstream Mach number M1 becomes large i.e., in the limit of strong shock
waves. In fact, from Eqs. (2.44) and (2.45) it is clear that
 
2
+1
lim
=
,
(2.50)
M1 1
1
and

1
.
(2.51)
2
Of course, before the Mach number gets large enough for these limits to be valid, real gas effects, e.g.,
ionization and/or dissociation of the gas molecules, may render the assumption of constant specific heats
invalid.
lim M2 =

M1

28

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

V2

V1

Figure 2.8: Flow through an oblique shock wave, oriented at the shock angle relative to the upstream
velocity vector. Since the tangential component of velocity is continuous across the shock, while the normal
component decreases, the flow is turned through the angle .
The normal shock relations of Eqs. (2.43), (2.44), (2.45), (2.46), and (2.47), are included in Tables 2.1 and 2.2
in Section 2.8 for a calorically perfect gas having ratio of specific heats = 1.4 (representative of air).

2.4

Oblique Shock Waves

Oblique shock waves are not normal to the upstream flow direction; rather the angle between the upstream
flow direction and the shock surface is the shock angle < /2, as illustrated in Fig. 2.8. Application of
the Momentum Theorem to a control volume of infinitesimal thickness spanning the shock shows that the
tangential component of velocity is the same on both sides of the shock, and the velocity component normal
to the shock surface changes in exactly the same way as that for a normal shock. In other words, if we view
the properties of the shock in a reference frame moving along the shock surface at a velocity equal to the
tangential component of velocity, the situation reduces to that of a normal shock. For the shock to exist,
the normal component of velocity upstream of the shock must be supersonic, and the normal component
downstream of the shock will be subsonic. Since the tangential component of velocity is the same on both
sides, the flow will be turned through some angle , as illustrated in the figure.
From the geometry illustrated in Fig. 2.8 it is easy to see that




Vn2
Vn2
tan ,
= tan1
= tan1
Vt2
Vn1
since Vt2 = Vt1 . This equation can be recast as
= tan

Mn2
Mn1

T2
tan
T1

where we have expressed the velocities in terms of Mach numbers and used the fact that the speed of sound is
proportional to the square root of the temperature. Using the normal shock relations Eqs. (2.45) and (2.46)
allows us to write this equation as
!

1
2
M

1
1
+
n
+1
1
= tan1
tan .
M2n1
Finally, since Mn1 = M1 sin , we can write
= tan

2 + ( 1)M21 sin2
( + 1)M21 sin cos

(2.52)

29

2.4. OBLIQUE SHOCK WAVES

90

80

Shock angle, ; degrees

70

60

50

40
M =1.2
M =1.3
M =1.5
M =1.7
M =2
M =2.5
M =3
M =3.5
M =4
M =5

30

20

10

10

15
20
25
30
Turning angle, ; degrees

35

40

45

Figure 2.9: Shock angle as a function of turning angle for oblique shocks in air (approximated as a calorically
perfect gas with = 1.4). Broken blue line shows locus of maximum turning angle for a given Mach number.
Broken red line represents dividing line separating flows for which the downstream flow is subsonic from those
for which downstream flow remains supersonic; points above the broken red line have subsonic downstream
flow, while those below it have supersonic downstream flow.

This equation allows us to determine the turning angle for given values of the free stream Mach number
M1 and the shock angle .
The relation between shock angle and turning angle given by Eq. (2.52) is shown in Fig. 2.9 for various free
stream Mach numbers; the ratio of specific heats is taken to be = 1.4. It is seen that there is a maximum
turning angle possible for a given upstream Mach number. For any turning angle less than this maximum,
there are two different shock angles that produce the same turning angle . Zero turning is achieved for
both, the case of a normal shock = /2 and for the limiting case of an infinitesimally weak shock inclined
at the Mach angle = sin1 (1/M).
The flow downstream of normal shock waves must be subsonic, while the flow downstream of infinitesimally
weak oblique shocks must remain supersonic. (If the tangential component of velocity, which does not change
across the shock, is large enough, the Mach number downstream of the shock can be supersonic even though
the normal component must be subsonic.) For a given Mach number, shock angles greater than that for
maximum turning angle are said to correspond to the strong branch of flow solutions, while shock angles
less than that for the maximum turning angle are said to correspond to the weak branch of flow solutions.
More simply, shocks on the strong branch are more like normal shocks while those on the weak branch are
more like Mach waves. As seen in Fig. 2.9, all strong shocks have subsonic downstream Mach numbers, and
most weak shocks have supersonic downstream Mach numbers but shocks on the weak branch very near
the maximum turning angle can have subsonic downstream flow.
The flow upstream of shock waves must be supersonic, but such supersonic flows can be achieved even when
the flight Mach number is subsonic (since the flow accelerates around the body, and especially above the
wing surface). Figure 2.10 dramatically visualizes the extensive shock patterns in the vicinity of an aircraft
in a low-altitude pass over San Francisco Bay at nearly the speed of sound. In particular, compare the shock
patterns at the downstream end of the supersonic zones surrounding the wing with the shock waves shown
in the higher Mach number flows past the airfoils shown in Fig. 2.14.

30

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

Figure 2.10: Photograph of an aircraft in flight at a Mach number only slightly less than unity. Shock waves
are made visible by diffraction of light through the sharp density gradients associated with them. They are
visible above the canopy, upstream of the engine inlets, and at the downstream end of the supersonic zones
above and below the wings. The background is blurred because the photographer was following the aircraft
with his camera; we might be able to estimate the speed of the aircraft if we knew the shutter speed and the
distances from the photographer to the aircraft and to the sailboats in the background.

2.5

Heat Addition

The flow in a gas turbine combustor is very complicated. Liquid fuel is injected as a spray that is then
atomized and evaporated, mixed with air, and burnt. The mixing of fuel and air, and the resulting exothermic
reaction take place in a turbulent flow, on length and time scales that make simulation extremely difficult.
Nevertheless, for simple parametric studies of engine performance, the net effect of the combustion can be
approximated as a simple addition of heat to the gas and it is this heating process of a compressible fluid
that we will characterize in this section.

2.5.1

Simple Heating the Rayleigh Process

We consider the addition of heat to the steady flow of a calorically perfect gas in a duct of constant crosssectional area, in the absence of viscous stresses; such a process is called simple heating or a Rayleigh process.
The differential forms of the continuity equation
dV
d
+
= 0,

(2.53)

dp + V dV = 0 ,

(2.54)

and the momentum equation


can be combined to eliminate the velocity V , giving
dp m
2

d
= 0,
2

(2.55)

where m
= V is the (constant) mass flow rate per unit cross-sectional area. This equation can be integrated
to give the purely thermodynamic relation
p+

m
2
= Const. .

(2.56)

31

Temperature, T

2.5. HEAT ADDITION

Entropy, s

Figure 2.11: Rayleigh line, illustrating points accessible by simple heating of a compressible flow in a duct
of constant cross-sectional area. Blue square indicates point of maximum static temperature, corresponding

to M = 1/ ; red circle indicates point of maximum entropy, corresponding to sonic flow (M = 1).
A plot of this equation in the T -s (or h-s) plane is called a Rayleigh line; a typical Rayleigh line is shown in
Fig. 2.11.
Since

dp
= 2 =V2,
d

at the point on the curve corresponding to maximum entropy,



dp
p
2
= a2 ,
=V =
d
s

(2.57)

(2.58)

so the flow is sonic at this point. The upper branch of the curve in Fig. 2.11 corresponds to subsonic flow,
and the lower branch corresponds to supersonic flow Since adding heat increases the entropy of the gas, this
must always drive the flow toward sonic conditions. In other words, adding heat to a subsonic flow will
increase its Mach number, while adding heat to a supersonic flow will decrease its Mach number.
At the point on the curve corresponding to maximum (static) temperature,

dp
a2
p
= RT =
=V2 =
,
(2.59)
d
T

so the Mach number at this point is M = 1/ . Note that, for Mach numbers in the range 1/ M 1,
the temperature decreases as heat is added. Remember, however, that the flow is accelerating toward M = 1,
so this simply means that the kinetic energy is increasing more rapidly than the rate at which heat is being
added.
In view of the preceding, the sonic point is again a convenient reference point for processes involving simple
heating. Equation 2.56 can be written as

which can be recast into the form

p + V 2 = Const. = p + V 2 ,

(2.60)

+1
p
=
.

p
1 + M2

(2.61)

32

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

2.5

3
Total Pressure
Total Temperature
Temperature
Pressure

2.5

Ratio to sonic value

Ratio to sonic value

Total Pressure
Total Temperature
Temperature
Pressure

1.5

1.5

0.5
0.5

0
0

0.1

0.2

0.3

0.4
0.5
0.6
Mach number, M_1

0.7

0.8

0.9

0
0

0.5

1.5
2
2.5
Mach number, M_1

3.5

Figure 2.12: Variation of flow properties for Rayleigh process; calorically perfect gas with = 1.4.
The density appearing in the continuity equation
V = a
can be eliminated in favor of the pressure p and temperature T using the equation of state of a perfect gas;
then, substitution of Eq. (2.61) results in an equation for the temperature ratio
( + 1)2 M2
T
=
.
T
(1 + M2 )2

(2.62)

The ratio of stagnation temperatures can be written


TT
TT T T
=
TT
T T TT


1 2
( + 1)2 M2
2
= 1+
M

2
2
2
(1 + M ) + 1
or
( + 1)M2 2 + ( 1)M2
TT
=
2
TT
(1 + M2 )

(2.63)

Finally, the ratio of stagnation pressures can be written


pT
pT p p
=

pT
p p pT
/(1)
/(1)


( + 1)
2
1 2

= 1+
2
1 + M2
+1
or
pT
+1
=

pT
1 + M2

2 + ( 1)M2
+1

/(1)

(2.64)

Equations (2.61) (2.64) are plotted for a gas having = 1.4 in Figs. 2.12. Data for the Rayleigh process
are also given in Tables 2.3 and 2.4 for = 1.4 and in Tables 2.5 and 2.6 for = 1.325 in Section 2.8.

33

2.6. CRITICAL MACH NUMBER

2.5.2

Heating at Constant Mach Number

As an alternative to heating in a constant-area duct, we also consider the addition of heat to a steady flow
at constant Mach number. Since we can write the mass flow rate m
as
r
r
p
TT
pT
(2.65)
m
= V A =
M A ,
R pT
T
TT
if the Mach number is held constant we must have
p
T A = Const. .
TT

(2.66)

Thus, the stagnation pressure ratio for constant Mach number heating between stations ()1 and ()2 can be
written as
s
pT2
TT2 A1
=
.
(2.67)
pT1
TT1 A2
Thus, in order to predict the loss in stagnation pressure we need to know how the cross-sectional area changes
as heat is added.
The continuity equation can be expressed for constant Mach number as
dV
dA
d
dT
dA
d
+
+
=
+
+
= 0,

V
A

2T
A

(2.68)

or, using the perfect gas law to eliminate the change in density in favor of pressure and temperature, we
have
dp
dT
dA

+
= 0.
(2.69)
p
2T
A
The Bernoulli equation for constant Mach number can be written as
dp
dT
dp V dV
+
=
+ M2
= 0,
p
p
p
2T

(2.70)

and eliminating the change in pressure between this and the continuity equations gives
1 + M2 dT
dA
=
.
A
2
T

(2.71)

Since the Mach number is constant, this can be integrated to give


A1
=
A2

TT1
TT2

(1+M2 )/2

(2.72)

Finally, combining Eqs. (2.67) and (2.72) gives


pT2
=
pT1

TT2
TT1

M2 /2

(2.73)

A comparison of the stagnation pressure loss for Rayleigh and constant-Mach heating is shown in Fig. 2.13.
For the Mach numbers shown, constant-Mach heating is seen to result in less total pressure loss for subsonic
flows, and greater total pressure loss for supersonic flows.

34

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

0.9

0.7
M =0.2

Stagnation pressure ratio, pt /pt

0.8

M =0.4
1

0.6

M =0.7
1

M =1.5
1

0.5

M =2
1

M =4
1

0.4

0.3

0.2

0.1

0.1

0.2

0.3

0.4
0.5
0.6
Total temperature ratio, Tt/Tt

0.7

0.8

0.9

Figure 2.13: Stagnation pressure ratio for heat addition in compressible flow. Calorically perfect gas with
specific heat ratio = 1.4. Solid lines represent heating at constant Mach number; dashed lines represent
heating at constant area (and their termination corresponds to thermal choking).

2.6

Critical Mach Number

The nature of the flow pattern past an airfoil as the free stream Mach number M is increased is illustrated
in Fig. 2.14. The flow here is that past the RAE 2822 airfoil at an angle of attack of = 3 , and the data
presented are based on numerical solutions of the inviscid Euler equations [3].
From the contours of constant Mach number shown in this figure, it is seen that the maximum Mach number
occurs on the airfoil surface, usually near the leading edge. As the free stream Mach number increases,
so does the maximum Mach number in the flow, and, for sufficiently large free stream Mach number the
maximum Mach number in the flow is supersonic. The free stream Mach number for which the maximum
Mach number in the flow becomes sonic is called the critical Mach number ; the critical Mach number for
this airfoil at this angle of attack is approximately Mcrit = 0.52.
Figure 2.15 shows the airfoil surface pressure distributions for several of these cases. The Mach number
M = 0.05 is sufficiently small that this case can be taken to correspond to essentially incompressible flow.
As long as the flow remains entirely subsonic (i.e., as long as the free stream Mach number is less than
the critical Mach number) the pressure distribution has a shape similar to that for incompressible flow. As
supersonic flow begins to develop, however, the pressure distribution changes significantly, and includes a
shock wave that terminates the supersonic pocket.
As long as the flow remains subsonic, the properties of flow past thin, two-dimensional airfoils can be related
to those of the incompressible flow past the same body. The Prandtl-Glauert similarity law states that the
pressure coefficient
p p
,
(2.74)
Cp 1
2
2 V
at any point on a thin airfoil for the compressible flow at a free stream Mach number M is related to the
pressure coefficient Cp 0 at the same point for incompressible flow according to
Cp = p

Cp 0
1 M2

(2.75)

35

2.6. CRITICAL MACH NUMBER

M = 0.20

M = 0.40

M = 0.50

M = 0.60

M = 0.70

M = 0.75

Figure 2.14: Contours of constant Mach number for inviscid flow past RAE 2822 airfoil at = 3.0 angle
of attack. Contour spacing is M = 0.05. Solutions to the Euler equations of inviscid flow are computed
numerically on 320 64 cell grids using Jameson-Caughey LU -SGS scheme [3].

36

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

Surface Pressure, Total Enthalpy, and Entropy Change Distributions


-2

Surface Pressure, Total Enthalpy, and Entropy Change Distributions


-2

Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)

-1.5

-1.5

-1
Pressure Coefficient

-1
Pressure Coefficient

Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)

-0.5

-0.5

0.5

0.5

1.5

1.5
0

0.2

0.4
0.6
Chordwise position, x/c

0.8

0.2

0.4
0.6
Chordwise position, x/c

M = 0.05

Surface Pressure, Total Enthalpy, and Entropy Change Distributions


-2

Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)

-1.5

Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)

-1.5

-1
Pressure Coefficient

-1
Pressure Coefficient

M = 0.50

Surface Pressure, Total Enthalpy, and Entropy Change Distributions


-2

0.8

-0.5

-0.5

0.5

0.5

1.5

1.5
0

0.2

0.4
0.6
Chordwise position, x/c

M = 0.60

0.8

0.2

0.4
0.6
Chordwise position, x/c

0.8

M = 0.70

Figure 2.15: Airfoil surface pressure coefficients for inviscid flow past RAE 2822 airfoil at = 3.0 angle
of attack. Solutions to the Euler equations of inviscid flow are computed numerically on 320 64 cell grids
using Jameson-Caughey LU -SGS scheme [3].

37

2.6. CRITICAL MACH NUMBER

Prandtl-Glauert Scaling for RAE 2822 Airfoil


Lift, numerical
Lift, Prandtl-Glauert
Drag, numerical

1.4

0.12
0.1

0.08

0.8

0.06

0.6

0.04

0.4

0.02

0.2

Drag coefficient

Lift coefficient

1.2

-0.02
0

0.2

0.4
0.6
Free stream Mach number

0.8

Figure 2.16: Lift and drag coefficients as functions of Mach number for inviscid flow past RAE 2822 airfoil
at = 3.0 angle of attack. Solutions to the Euler equations of inviscid flow are computed numerically on
320 64 cell grids using Jameson-Caughey LU -SGS scheme [3]; lift coefficient is compared with results of
the Prandtl-Glauert similarity law.
Since the lift coefficient Cl is equal to the integral of the difference in pressure coefficients on the lower and
upper surfaces of the airfoil, a similar relation holds for it
Cl = p

Cl 0
,
1 M2

(2.76)

where Cl 0 is the lift coefficient in the corresponding incompressible flow. A comparison of the lift coefficient computed using the Prandtl-Glauert rule with that computed numerically for the airfoil of the earlier
Figs. 2.14 and 2.15 is shown in Fig. 2.16. The Prandtl-Glauert rule is seen to predict the increase in lift coefficient with increasing Mach number quite well up to approximately the critical Mach number (of Mcrit = 0.52
for this case). The computed drag coefficient is also plotted in the figure, showing a rapid rise as the size of
the supersonic zone and the strength and extent of the shock wave increase for Mach numbers larger than
about M = 0.65.
A similar relation also holds in supersonic flow, giving the pressure coefficient at any point on the airfoil as
Cp = p

Cp 2
M2 1

(2.77)

where Cp 2 represents the pressure coefficient at the same point for the flow at a free stream Mach number

M = 2. Both the subsonic and supersonic Prandtl-Glauert rules can be derived by studying the way in
which the equations describing small perturbations from the free stream conditions must scale with changes
in the Mach number. For a description of this procedure, and for the more general scaling laws that apply
to three-dimensional flows, see, e.g., the aerodynamics texts [1] or [2].
As was seen above, the flow past a body begins to behave in a qualitatively different fashion from an
incompressible one when the maximum local Mach number in the flow exceeds unity, and a pocket of
supersonic flow begins to form. This supersonic pocket is almost always terminated by a shock wave, and
the losses associated with the shock wave appear as drag on the airfoil. As the strength and extent of the
shock wave grow with increasing free stream Mach number, the drag begins to increase dramatically. The
free stream Mach number for which the derivative of the drag coefficient with respect to Mach number
reaches
CD
= 0.10
(2.78)
M

38

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

is called the drag divergence Mach number .


The critical Mach number is necessarily a conservative estimate for the drag divergence Mach number. And
the critical Mach number is easily determined by finding the Mach number at which the Prandtl-Glauert
rule predicts the minimum pressure coefficient will become equal to its sonic value.
For an isentropic flow, the pressure coefficient can be expressed in terms of the free stream and local Mach
numbers using


2
p/p 1
p pT
Cp = 2 =

1 ,
(2.79)
M2 pT p
2p V
which, using the isentropic relation, Eq. (2.21), can be written

!/(1)
1
2 1 + 2 M2
1 ,
Cp =
2
M2
1 + 1
M
2

(2.80)

where M is the local Mach number, corresponding to the point in the flow where the pressure is p, and M
is the Mach number in the free stream where the pressure is p .
The value of the critical pressure coefficient Cp at a point where the local Mach number is unity is then
given by
"
#
/(1)
2
1 2
2

+
M
1 .
(2.81)
Cp =
M2
+1 +1
A plot of the critical pressure coefficient as a function of free stream Mach number is shown in Fig. 2.17, along
with plots of the Prandtl-Glauert scaling given by Eq. (2.75) for several values of incompressible pressure
coefficient of Cp 0 . Since the maximum Mach number in isentropic flow will occur at the point where the
pressure is the lowest and, according to the Prandtl-Glauert rule the minimum pressure will always occur at
the same point, the point on the airfoil at which sonic flow will first occur is the point corresponding to the
minimum pressure point in the incompressible flow past the body. Thus, the Mach number at which these
two curves intersect (seen in the figure to be approximately M = 0.575, .643, and .747, respectively for the
several values of Cp 0 plotted) will be the critical Mach number for an airfoil having the minimum pressure
coefficient specified in incompressible flow. The critical Mach numbers for other values of Cp 0 can be found
by a similar procedure.

2.7

Supersonic Airfoil Theory

Linearized supersonic flow theory relates the pressure coefficient at any point on a two-dimensional airfoil to
the local slope of the surface relative to the free stream direction according to
Cp = p

2
,
M2 1

(2.82)

where the angle is positive when the surface normal has a downstream component. This simple result can
be used to investigate the properties of airfoils in supersonic flow.
We represent the coordinates of the upper surface yu and the lower surface yl of the airfoil as
x
yu
+ fc (x/c) + ft (x/c)
=
c
c

yl
x
+ fc (x/c) ft (x/c) ,
=
c
c

(2.83)

39

2.7. SUPERSONIC AIRFOIL THEORY

3.5

*
p

= 1.2

= .80

2.5

= .40

Pressure Coefficient, C

1.5

0.5

0.1

0.2

0.3

0.4
0.5
0.6
Mach number, M

0.7

0.8

0.9

Figure 2.17: Critical pressure coefficient and Prandtl-Glauert scaled pressure coefficients (for Cp 0 =
0.4, .8, 1.2) as functions of free stream Mach number M .
where the functions fc and ft represent the contributions of camber and thickness, respectively. Since the
angle of attack is represented separately, we have
fc (0) = fc (1) = 0
ft (0) = ft (1) = 0 .

(2.84)

The lift coefficient is given by the integral


Cl =



Cp l Cp u dx/c ,

which, using the linearized supersonic flow result of Eq. (2.82), can be written

 
Z 1
dy
2
dy
Cl = p
dx/c .
+
dx l
dx u
M2 1 0

(2.85)

(2.86)

By virtue of the conditions imposed by Eqs. (2.84), the above integral reduces to
Cl = p

2
4
[ ] = p
.
2
M 1
M2 1

(2.87)

It is seen that, for supersonic flow the lift coefficient is completely independent of the camber and thickness
distributions of the airfoil, and depends only on the angle of attack.
The drag coefficient is given by the integral


 

Z 1
dy
dy
dx/c
+ Cp u
Cp l
Cd =
dx l
dx u
0
"
#
2 
2
Z 1 
dy
dy
2
dx/c .
+
=p
dx l
dx u
M2 1 0

Substituting for the airfoil geometry using Eqs. (2.83), the drag coefficient can be expressed as


Z 1

4
2
2
2 +
fc + ft dx/c .
Cd = p
M2 1
0

(2.88)

(2.89)

40

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

Thus, while thickness and camber contribute nothing to the lift coefficient, they do contribute to the drag
coefficient. Thickness is, of course, required for structural reasons, so optimal supersonic wings require a
careful compromise between thicker wings that will generally weigh less and thinner wings that will have
less drag. Camber serves no useful purpose in supersonic flow, so wings optimized for supersonic flight will
have little, or no, camber.
The integrals in Eq. (2.89) can be evaluated for particular camber line shapes and thickness distributions.
For example, the circular-arc airfoil having thickness ratio is defined by
x
x
1
ft = 2
c
c
so


x 2
2
ft = 4 2 1 2
c

and the drag coefficient for this thickness distribution is given by


16 2
Cd = p
.
3 M2 1

(2.90)

It can be shown that the minimum drag coefficient for an airfoil of given thickness ratio is

for the double-diamond shape defined as

Cd = p
ft = min

4 2
M2 1

(2.91)

x h
x i
, 1
c
c

Finally, note that since the lift coefficient for any airfoil shape is the same as that for a flat plate at the same
angle of attack, the lift will be distributed uniformly along the chord line, so the centroid of the incremental
lift due to an infinitesimal change in angle of attack will be at mid-chord. This means that the aerodynamic
center of a supersonic airfoil will be at the mid-chord station (rather than at the quarter-chord station, as
in low-speed flow).

2.7.1

Prandtl-Meyer Flow

Equation (2.82) can be developed in a variety of ways. The most direct way is from a linearization of the
equations of inviscid, supersonic flow, but the result can also be developed from a description of PrandtlMeyer flow. Inviscid, supersonic flow of a uniform stream past a sharp corner joining two straight walls must
consist of two uniform regions separated by an expansion fan, as sketched in Fig. 2.18 (a). Each ray of the
fan is inclined relative to the local flow direction at the Mach angle corresponding to the Mach number of
the flow on that ray which is uniform along the ray. Thus, the first and last rays of the fan are inclined
at the initial and final Mach angles relative to the upstream and downstream walls, respectively. We can
develop the equations describing the relationship between flow angle and pressure (or velocity) by considering
a differential element of the turn or, equivalently, the flow past a corner of infinitesimal turning angle, as
sketched in Fig. 2.18 (b).
Applying the conservation of mass to the control volume indicated by the broken lines in Fig. 2.18 gives for
the change in the component of velocity normal to the wave
dVn = Vn

d
.

(2.92)

41

2.7. SUPERSONIC AIRFOIL THEORY

V1

M1

V2

M2

(a)

(b)

Figure 2.18: Prandtl-Meyer expansion fan. (a) Finite turning angle; (b) Infinitesimal turning angle.
Applying conservation of momentum to the same control volume shows that the tangential component of
velocity Vt cannot change across the wave, and the change in normal component of velocity Vn is related to
the pressure change dp by
dp
.
(2.93)
dVn =
Vn
Combining these equations to eliminate dVn gives
dp
= Vn2 ,
d

(2.94)

which confirms the fact that the ray is a Mach wave (i.e., that Vn = a, the local speed of sound).
To relate the change in velocity to the turning angle we note that, by virtue of the fact that the tangential
component of velocity is the same on both sides of the wave
Vt = V cos = (V + dV ) cos ( + d) ,

(2.95)

where = sin1 (1/M) is the Mach angle. In the limit as infinitesimals vanish, this gives
d = cot

p
dV
dV
= M2 1
,
V
V

and replacing the right-hand side using the Bernoulli equation gives

M2 1 dp
,
d =
M2
p

(2.96)

(2.97)

which is an exact formula relating pressure changes to turning angle for this flow.
If the isentropic relation
M dM
dp
=
2
p
1 + 1
2 M

(2.98)

is substituted into Eq. (2.97), it can be integrated to give


=

+1
1

1/2

tan

"

 #1/2
p
( 1) M2 1
tan1 M2 1 ,
+1

(2.99)

where the constant of integration has been chosen so that = 0 at sonic (M = 1) conditions. This is called
the Prandtl-Meyer function, and relates the Mach number to the turning angle in the supersonic flow past
a sharp corner.

42

CHAPTER 2.

SUMMARY OF RESULTS FROM COMPRESSIBLE FLOW THEORY

For the present purpose, however, we are just interested in the linearized form of this function, corresponding
to turning of a supersonic flow through small angles. For this purpose, we use the differential form of
Eq. (2.79)
2 dp
dCp =
(2.100)
M2 p
to express the pressure change in Eq. (2.97) in terms of the pressure coefficient, giving
2 d
,
dCp =
M2 1
which integrates to
Cp =

2
,
M2 1

which is equivalent to Eq. (2.82) for small turning angles relative to the free stream direction.

2.7.2

Turning by Weak Shocks

It is of interest to see what is the linear approximation to the shock turning relation of Eq. (2.52). This will
provide an approximation suitable for use when the turning angle is small. If we represent
M2n1 = 1 +
then in the limit of small the shock angle will be very nearly equal to the Mach angle ( sin1 1/M)
and Eq. (2.52) can be approximated as
p
2 M21 1

(2.101)
=
( + 1)M21
The parameter can be related to the pressure coefficient
p2 p 1
2
Cp = 1 2 =
M21
2 V1


p2
1
p1

using the normal shock relation, Eq. (2.43), which is equivalent to




+ 1 p2
1 .
=
2
p1
Thus, we have
Cp =

4
,
( + 1)M21

or
Cp = p

2
M21 1

(2.102)

(2.103)

(2.104)

(2.105)

This equation is equivalent to Eq. (2.82), except for the sign. The signs are different only because the
positive sense of the turning angle is defined oppositely in the two cases: for the Prandtl-Meyer flow,
positive corresponds to an expanding flow in which the outward normal to the surface has a component in
the downstream direction, while for the oblique shock positive corresponds to a compressive flow in which
the surface normal has an upstream component.

Bibliography
[1] Holt Ashley & Marten Landahl, Aerodynamics of Wings and Bodies, Addison-Wesley, 1965.
[2] John J. Bertin, Aerodynamics for Engineers, Fourth Edition, Prentice Hall, 2002.
[3] A. Jameson & D. A. Caughey, How Many Steps are Required to Solve the Euler Equations of Steady,
Compressible Flow: In Search of a Fast Solution Algorithm, AIAA Paper 2001-2673, AIAA 15th
Computational Fluid Dynamics Conference, June 11-14, Anaheim, California.
[4] H. W. Liepmann & A. Roshko, Elements of Gasdynamics, John Wiley & Sons, 1957.
[5] Jack Mattingly, Elements of Propulsion: Gas Turbines and Rockets, American Institute of
Aeronautics and Astronautics, 2006.
[6] Asher H. Shapiro, The Dynamics and Thermodynamics of Compressible Fluid Flow, Volume
I, Ronald Press, 1953.

2.8

Compressible Flow Tables

Tables of compressible flow properties for adiabatic flow (both isentropic and including normal shock waves)
are included on the following pages for calorically perfect air, having = 1.4.

43

44

BIBLIOGRAPHY
M
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
0.18
0.20
0.22
0.24
0.26
0.28
0.30
0.32
0.34
0.36
0.38
0.40
0.42
0.44
0.46
0.48
0.50
0.52
0.54
0.56
0.58
0.60
0.62
0.64
0.66
0.68
0.70
0.72
0.74
0.76
0.78
0.80
0.82
0.84
0.86
0.88
0.90
0.92
0.94
0.96
0.98
1.00

T
TT

p
pT

A
A

1.000000
0.999920
0.999680
0.999281
0.998722
0.998004
0.997128
0.996095
0.994906
0.993562
0.992063
0.990413
0.988611
0.986660
0.984562
0.982318
0.979931
0.977402
0.974735
0.971931
0.968992
0.965922
0.962723
0.959398
0.955950
0.952381
0.948695
0.944894
0.940982
0.936961
0.932836
0.928609
0.924283
0.919862
0.915349
0.910747
0.906060
0.901291
0.896443
0.891520
0.886525
0.881461
0.876332
0.871141
0.865891
0.860585
0.855227
0.849820
0.844366
0.838870
0.833333

1.000000
0.999720
0.998881
0.997484
0.995533
0.993031
0.989985
0.986400
0.982285
0.977647
0.972497
0.966845
0.960703
0.954085
0.947002
0.939470
0.931503
0.923118
0.914330
0.905156
0.895614
0.885722
0.875498
0.864960
0.854128
0.843019
0.831654
0.820050
0.808228
0.796206
0.784004
0.771639
0.759131
0.746498
0.733758
0.720928
0.708025
0.695068
0.682070
0.669050
0.656022
0.643000
0.630000
0.617034
0.604117
0.591260
0.578476
0.565775
0.553170
0.540669
0.528282

1.000000
0.999800
0.999200
0.998202
0.996807
0.995017
0.992836
0.990267
0.987314
0.983982
0.980277
0.976204
0.971771
0.966984
0.961851
0.956380
0.950580
0.944460
0.938029
0.931297
0.924274
0.916971
0.909398
0.901566
0.893486
0.885170
0.876629
0.867876
0.858920
0.849775
0.840452
0.830963
0.821319
0.811533
0.801616
0.791579
0.781434
0.771191
0.760863
0.750460
0.739992
0.729471
0.718905
0.708306
0.697683
0.687044
0.676400
0.665759
0.655130
0.644520
0.633938

28.942130
14.481486
9.665910
7.261610
5.821829
4.864318
4.182400
3.672739
3.277926
2.963520
2.707602
2.495562
2.317287
2.165554
2.035065
1.921851
1.822876
1.735778
1.658696
1.590140
1.528905
1.474005
1.424629
1.380097
1.339844
1.303388
1.270321
1.240294
1.213007
1.188200
1.165646
1.145148
1.126535
1.109655
1.094373
1.080571
1.068144
1.056999
1.047053
1.038230
1.030464
1.023696
1.017871
1.012941
1.008863
1.005597
1.003108
1.001365
1.000337
1.000000

V /a
0.000000
0.021908
0.043811
0.065703
0.087580
0.109435
0.131265
0.153063
0.174824
0.196544
0.218218
0.239840
0.261405
0.282910
0.304348
0.325715
0.347007
0.368219
0.389347
0.410385
0.431331
0.452180
0.472927
0.493569
0.514102
0.534522
0.554826
0.575011
0.595072
0.615006
0.634811
0.654483
0.674020
0.693419
0.712677
0.731792
0.750761
0.769582
0.788253
0.806772
0.825137
0.843347
0.861399
0.879292
0.897026
0.914598
0.932007
0.949253
0.966334
0.983250
1.000000

MFP
0.000000
0.023659
0.047283
0.070840
0.094295
0.117614
0.140766
0.163717
0.186436
0.208892
0.231053
0.252892
0.274380
0.295488
0.316192
0.336467
0.356287
0.375633
0.394481
0.412813
0.430611
0.447857
0.464538
0.480639
0.496147
0.511053
0.525347
0.539022
0.552072
0.564491
0.576277
0.587427
0.597941
0.607821
0.617067
0.625684
0.633676
0.641048
0.647807
0.653961
0.659518
0.664488
0.668882
0.672710
0.675984
0.678716
0.680920
0.682610
0.683798
0.684501
0.684731

Table 2.1: Subsonic, adiabatic flow of ideal gas with = 1.4.

45

2.8. COMPRESSIBLE FLOW TABLES


M
1.00
1.10
1.20
1.30
1.40
1.50
1.60
1.70
1.80
1.90
2.00
2.10
2.20
2.30
2.40
2.50
2.60
2.70
2.80
2.90
3.00
3.10
3.20
3.30
3.40
3.50
3.60
3.70
3.80
3.90
4.00
4.10
4.20
4.30
4.40
4.50
4.60
4.70
4.80
4.90
5.00
5.10
5.20
5.30
5.40
5.50
5.60
5.70
5.80
5.90
6.00

T
TT

p
pT

A
A

0.83333
0.80515
0.77640
0.74738
0.71839
0.68966
0.66138
0.63371
0.60680
0.58072
0.55556
0.53135
0.50813
0.48591
0.46468
0.44444
0.42517
0.40683
0.38941
0.37286
0.35714
0.34223
0.32808
0.31466
0.30193
0.28986
0.27840
0.26752
0.25720
0.24740
0.23810
0.22925
0.22085
0.21286
0.20525
0.19802
0.19113
0.18457
0.17832
0.17235
0.16667
0.16124
0.15605
0.15110
0.14637
0.14184
0.13751
0.13337
0.12940
0.12560
0.12195

0.52828
0.46835
0.41238
0.36091
0.31424
0.27240
0.23527
0.20259
0.17404
0.14924
0.12780
0.10935
0.09352
0.07997
0.06840
0.05853
0.05012
0.04295
0.03685
0.03165
0.02722
0.02345
0.02023
0.01748
0.01512
0.01311
0.01138
0.00990
0.00863
0.00753
0.00659
0.00577
0.00506
0.00445
0.00392
0.00346
0.00305
0.00270
0.00239
0.00213
0.00189
0.00168
0.00150
0.00134
0.00120
0.00107
0.00096
0.00087
0.00078
0.00070
0.00063

1.00000
1.00793
1.03044
1.06630
1.11493
1.17617
1.25023
1.33761
1.43898
1.55526
1.68750
1.83694
2.00497
2.19313
2.40310
2.63672
2.89598
3.18301
3.50012
3.84977
4.23457
4.65731
5.12096
5.62865
6.18370
6.78962
7.45011
8.16907
8.95059
9.79897
10.71875
11.71465
12.79164
13.95490
15.20987
16.56219
18.01779
19.58283
21.26371
23.06712
25.00000
27.06957
29.28333
31.64905
34.17481
36.86896
39.74018
42.79743
46.05000
49.50747
53.17978

V /a
1.00000
1.08124
1.15828
1.23114
1.29987
1.36458
1.42539
1.48247
1.53598
1.58609
1.63299
1.67687
1.71791
1.75629
1.79218
1.82574
1.85714
1.88653
1.91404
1.93981
1.96396
1.98661
2.00786
2.02781
2.04656
2.06419
2.08077
2.09639
2.11111
2.12499
2.13809
2.15046
2.16215
2.17321
2.18368
2.19360
2.20300
2.21192
2.22038
2.22842
2.23607
2.24334
2.25026
2.25685
2.26313
2.26913
2.27484
2.28030
2.28552
2.29051
2.29528

MFP
0.68473
0.67935
0.66450
0.64215
0.61415
0.58217
0.54768
0.51191
0.47584
0.44027
0.40577
0.37276
0.34152
0.31222
0.28494
0.25969
0.23644
0.21512
0.19563
0.17786
0.16170
0.14702
0.13371
0.12165
0.11073
0.10085
0.09191
0.08382
0.07650
0.06988
0.06388
0.05845
0.05353
0.04907
0.04502
0.04134
0.03800
0.03497
0.03220
0.02968
0.02739
0.02530
0.02338
0.02164
0.02004
0.01857
0.01723
0.01600
0.01487
0.01383
0.01288

M2
1.00000
0.91177
0.84217
0.78596
0.73971
0.70109
0.66844
0.64054
0.61650
0.59562
0.57735
0.56128
0.54706
0.53441
0.52312
0.51299
0.50387
0.49563
0.48817
0.48138
0.47519
0.46953
0.46435
0.45959
0.45520
0.45115
0.44741
0.44395
0.44073
0.43774
0.43496
0.43236
0.42994
0.42767
0.42554
0.42355
0.42168
0.41992
0.41826
0.41670
0.41523
0.41384
0.41252
0.41127
0.41009
0.40897
0.40791
0.40690
0.40594
0.40503
0.40416

p2
p1

2
1

T2
T1

1.00000
1.24500
1.51333
1.80500
2.12000
2.45833
2.82000
3.20500
3.61333
4.04500
4.50000
4.97833
5.48000
6.00500
6.55333
7.12500
7.72000
8.33833
8.98000
9.64500
10.33333
11.04500
11.78000
12.53833
13.32000
14.12500
14.95333
15.80500
16.68000
17.57833
18.50000
19.44500
20.41333
21.40500
22.42000
23.45833
24.52000
25.60500
26.71333
27.84500
29.00000
30.17833
31.38000
32.60500
33.85333
35.12500
36.42000
37.73833
39.08000
40.44500
41.83333

1.00000
1.16908
1.34161
1.51570
1.68966
1.86207
2.03175
2.19772
2.35922
2.51568
2.66667
2.81190
2.95122
3.08455
3.21190
3.33333
3.44898
3.55899
3.66355
3.76286
3.85714
3.94661
4.03150
4.11202
4.18841
4.26087
4.32962
4.39486
4.45679
4.51559
4.57143
4.62448
4.67491
4.72286
4.76847
4.81188
4.85321
4.89258
4.93010
4.96587
5.00000
5.03257
5.06367
5.09338
5.12178
5.14894
5.17492
5.19979
5.22360
5.24642
5.26829

1.00000
1.06494
1.12799
1.19087
1.25469
1.32022
1.38797
1.45833
1.53158
1.60792
1.68750
1.77045
1.85686
1.94680
2.04033
2.13750
2.23834
2.34289
2.45117
2.56321
2.67901
2.79860
2.92199
3.04919
3.18021
3.31505
3.45373
3.59624
3.74260
3.89281
4.04687
4.20479
4.36657
4.53221
4.70171
4.87509
5.05233
5.23343
5.41842
5.60727
5.80000
5.99660
6.19709
6.40144
6.60968
6.82180
7.03779
7.25767
7.48143
7.70907
7.94059

Table 2.2: Supersonic, adiabatic flow of ideal gas with = 1.4.

pT2
pT1

1.00000
0.99893
0.99280
0.97937
0.95819
0.92979
0.89520
0.85572
0.81268
0.76736
0.72087
0.67420
0.62814
0.58329
0.54014
0.49901
0.46012
0.42359
0.38946
0.35773
0.32834
0.30121
0.27623
0.25328
0.23223
0.21295
0.19531
0.17919
0.16447
0.15103
0.13876
0.12756
0.11733
0.10800
0.09948
0.09170
0.08459
0.07809
0.07214
0.06670
0.06172
0.05715
0.05297
0.04913
0.04560
0.04236
0.03938
0.03664
0.03412
0.03179
0.02965

46

BIBLIOGRAPHY
M
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
0.18
0.20
0.22
0.24
0.26
0.28
0.30
0.32
0.34
0.36
0.38
0.40
0.42
0.44
0.46
0.48
0.50
0.52
0.54
0.56
0.58
0.60
0.62
0.64
0.66
0.68
0.70
0.72
0.74
0.76
0.78
0.80
0.82
0.84
0.86
0.88
0.90
0.92
0.94
0.96
0.98
1.00

TT
TT

0.000000
0.001918
0.007648
0.017119
0.030215
0.046777
0.066606
0.089471
0.115110
0.143238
0.173554
0.205742
0.239484
0.274459
0.310353
0.346860
0.383689
0.420565
0.457232
0.493456
0.529027
0.563758
0.597485
0.630068
0.661390
0.691358
0.719897
0.746952
0.772486
0.796478
0.818923
0.839825
0.859203
0.877084
0.893502
0.908499
0.922122
0.934423
0.945456
0.955279
0.963948
0.971524
0.978066
0.983633
0.988283
0.992073
0.995058
0.997293
0.998828
0.999715
1.000000

T
T

0.000000
0.002301
0.009175
0.020529
0.036212
0.056020
0.079698
0.106946
0.137429
0.170779
0.206612
0.244523
0.284108
0.324957
0.366674
0.408873
0.451187
0.493273
0.534816
0.575526
0.615148
0.653456
0.690255
0.725383
0.758707
0.790123
0.819554
0.846948
0.872274
0.895523
0.916704
0.935843
0.952976
0.968155
0.981439
0.992895
1.002598
1.010624
1.017057
1.021980
1.025477
1.027633
1.028533
1.028260
1.026894
1.024516
1.021201
1.017023
1.012052
1.006357
1.000000

pT
p
T

1.267876
1.267522
1.266460
1.264700
1.262256
1.259146
1.255394
1.251029
1.246083
1.240592
1.234596
1.228136
1.221255
1.214000
1.206416
1.198549
1.190446
1.182153
1.173714
1.165175
1.156577
1.147960
1.139364
1.130825
1.122377
1.114053
1.105882
1.097892
1.090109
1.082556
1.075253
1.068221
1.061475
1.055031
1.048904
1.043104
1.037642
1.032528
1.027769
1.023372
1.019343
1.015687
1.012407
1.009507
1.006989
1.004856
1.003109
1.001749
1.000778
1.000194
1.000000

p
p

2.400000
2.398657
2.394636
2.387965
2.378687
2.366864
2.352572
2.335903
2.316960
2.295860
2.272727
2.247696
2.220906
2.192502
2.162630
2.131439
2.099076
2.065689
2.031419
1.996406
1.960784
1.924681
1.888218
1.851509
1.814662
1.777778
1.740947
1.704255
1.667779
1.631588
1.595745
1.560306
1.525320
1.490831
1.456876
1.423488
1.390692
1.358511
1.326964
1.296064
1.265823
1.236247
1.207341
1.179106
1.151543
1.124649
1.098418
1.072846
1.047925
1.023646
1.000000

Table 2.3: Constant-area, subsonic flow of ideal gas with heat addition for = 1.4.

47

2.8. COMPRESSIBLE FLOW TABLES


M
1.00
1.10
1.20
1.30
1.40
1.50
1.60
1.70
1.80
1.90
2.00
2.10
2.20
2.30
2.40
2.50
2.60
2.70
2.80
2.90
3.00
3.10
3.20
3.30
3.40
3.50
3.60
3.70
3.80
3.90
4.00
4.10
4.20
4.30
4.40
4.50
4.60
4.70
4.80
4.90
5.00
5.10
5.20
5.30
5.40
5.50
5.60
5.70
5.80
5.90
6.00

TT
TT

1.000000
0.993924
0.978717
0.957979
0.934254
0.909276
0.884186
0.859709
0.836279
0.814136
0.793388
0.774064
0.756135
0.739543
0.724213
0.710059
0.696995
0.684935
0.673796
0.663502
0.653979
0.645162
0.636989
0.629405
0.622359
0.615805
0.609701
0.604010
0.598698
0.593732
0.589086
0.584733
0.580651
0.576818
0.573215
0.569825
0.566632
0.563621
0.560779
0.558094
0.555556
0.553153
0.550877
0.548718
0.546670
0.544725
0.542877
0.541118
0.539444
0.537849
0.536329

T
T

1.000000
0.960313
0.911848
0.859174
0.805391
0.752504
0.701735
0.653771
0.608941
0.567342
0.528926
0.493558
0.461058
0.431220
0.403836
0.378698
0.355610
0.334387
0.314858
0.296869
0.280277
0.264954
0.250783
0.237661
0.225492
0.214193
0.203686
0.193904
0.184783
0.176269
0.168310
0.160862
0.153883
0.147335
0.141186
0.135404
0.129961
0.124833
0.119995
0.115428
0.111111
0.107027
0.103160
0.099496
0.096019
0.092719
0.089584
0.086602
0.083765
0.081062
0.078487

pT
p
T

1.000000
1.004858
1.019415
1.043660
1.077652
1.121545
1.175611
1.240235
1.315925
1.403300
1.503096
1.616159
1.743446
1.886020
2.045055
2.221831
2.417741
2.634285
2.873080
3.135853
3.424452
3.740844
4.087118
4.465489
4.878303
5.328035
5.817298
6.348844
6.925566
7.550504
8.226849
8.957944
9.747289
10.598545
11.515538
12.502262
13.562884
14.701745
15.923366
17.232454
18.633900
20.132787
21.734395
23.444199
25.267881
27.211325
29.280629
31.482103
33.822275
36.307897
38.945945

p
p

1.000000
0.890869
0.795756
0.713012
0.641026
0.578313
0.523560
0.475624
0.433526
0.396432
0.363636
0.334541
0.308642
0.285510
0.264784
0.246154
0.229358
0.214171
0.200401
0.187882
0.176471
0.166044
0.156495
0.147729
0.139665
0.132231
0.125366
0.119012
0.113122
0.107652
0.102564
0.097823
0.093400
0.089266
0.085397
0.081772
0.078370
0.075174
0.072167
0.069336
0.066667
0.064147
0.061767
0.059515
0.057383
0.055363
0.053447
0.051628
0.049900
0.048257
0.046693

Table 2.4: Constant-area, supersonic flow of ideal gas with heat addition for = 1.4.

48

BIBLIOGRAPHY
M
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
0.18
0.20
0.22
0.24
0.26
0.28
0.30
0.32
0.34
0.36
0.38
0.40
0.42
0.44
0.46
0.48
0.50
0.52
0.54
0.56
0.58
0.60
0.62
0.64
0.66
0.68
0.70
0.72
0.74
0.76
0.78
0.80
0.82
0.84
0.86
0.88
0.90
0.92
0.94
0.96
0.98
1.00

TT
TT

0.000000
0.001858
0.007410
0.016591
0.029292
0.045365
0.064627
0.086860
0.111821
0.139241
0.168838
0.200314
0.233367
0.267691
0.302986
0.338959
0.375326
0.411819
0.448189
0.484205
0.519655
0.554352
0.588128
0.620840
0.652366
0.682603
0.711470
0.738905
0.764862
0.789313
0.812242
0.833648
0.853540
0.871939
0.888872
0.904376
0.918491
0.931265
0.942747
0.952990
0.962051
0.969985
0.976850
0.982703
0.987602
0.991601
0.994758
0.997125
0.998754
0.999696
1.000000

T
T

0.000000
0.002160
0.008612
0.019276
0.034017
0.052652
0.074953
0.100654
0.129453
0.161020
0.195007
0.231048
0.268773
0.307810
0.347791
0.388360
0.429175
0.469913
0.510274
0.549983
0.588790
0.626476
0.662846
0.697735
0.731006
0.762547
0.792271
0.820116
0.846038
0.870017
0.892046
0.912139
0.930318
0.946622
0.961097
0.973798
0.984788
0.994132
1.001904
1.008178
1.013029
1.016536
1.018775
1.019825
1.019760
1.018656
1.016585
1.013617
1.009821
1.005261
1.000000

pT
p
T

1.258406
1.258073
1.257076
1.255422
1.253124
1.250199
1.246668
1.242558
1.237896
1.232717
1.227055
1.220949
1.214437
1.207562
1.200366
1.192892
1.185182
1.177279
1.169226
1.161063
1.152830
1.144566
1.136307
1.128090
1.119946
1.111907
1.104003
1.096261
1.088705
1.081359
1.074244
1.067381
1.060786
1.054475
1.048463
1.042763
1.037386
1.032342
1.027641
1.023290
1.019296
1.015666
1.012404
1.009515
1.007003
1.004871
1.003122
1.001759
1.000783
1.000196
1.000000

p
p

2.325000
2.323768
2.320081
2.313962
2.305450
2.294597
2.281470
2.266148
2.248723
2.229296
2.207977
2.184883
2.160138
2.133869
2.106207
2.077284
2.047232
2.016181
1.984262
1.951600
1.918317
1.884529
1.850349
1.815881
1.781227
1.746479
1.711724
1.677041
1.642506
1.608184
1.574137
1.540419
1.507078
1.474159
1.441700
1.409732
1.378284
1.347381
1.317042
1.287283
1.258117
1.229554
1.201600
1.174260
1.147536
1.121428
1.095933
1.071049
1.046769
1.023089
1.000000

Table 2.5: Constant-area, subsonic flow of ideal gas with heat addition for = 1.325.

49

2.8. COMPRESSIBLE FLOW TABLES


M
1.00
1.10
1.20
1.30
1.40
1.50
1.60
1.70
1.80
1.90
2.00
2.10
2.20
2.30
2.40
2.50
2.60
2.70
2.80
2.90
3.00
3.10
3.20
3.30
3.40
3.50
3.60
3.70
3.80
3.90
4.00
4.10
4.20
4.30
4.40
4.50
4.60
4.70
4.80
4.90
5.00
5.10
5.20
5.30
5.40
5.50
5.60
5.70
5.80
5.90
6.00

TT
TT

1.000000
0.993493
0.977106
0.954626
0.928770
0.901422
0.873839
0.846833
0.820901
0.796325
0.773243
0.751696
0.731667
0.713100
0.695918
0.680033
0.665352
0.651785
0.639241
0.627637
0.616894
0.606939
0.597706
0.589131
0.581161
0.573743
0.566831
0.560384
0.554363
0.548733
0.543462
0.538523
0.533890
0.529538
0.525446
0.521595
0.517966
0.514544
0.511314
0.508261
0.505374
0.502641
0.500051
0.497595
0.495265
0.493051
0.490946
0.488944
0.487038
0.485222
0.483491

T
T

1.000000
0.965160
0.920491
0.870650
0.818882
0.767343
0.717400
0.669861
0.625154
0.583457
0.544785
0.509049
0.476106
0.445777
0.417874
0.392205
0.368583
0.346833
0.326789
0.308299
0.291224
0.275437
0.260823
0.247277
0.234705
0.223022
0.212151
0.202022
0.192573
0.183747
0.175493
0.167764
0.160519
0.153719
0.147330
0.141321
0.135662
0.130328
0.125296
0.120542
0.116049
0.111797
0.107770
0.103952
0.100330
0.096891
0.093623
0.090515
0.087556
0.084738
0.082052

pT
p
T

1.000000
1.004926
1.019791
1.044750
1.080035
1.125997
1.183128
1.252072
1.333634
1.428786
1.538671
1.664609
1.808106
1.970859
2.154768
2.361943
2.594720
2.855668
3.147608
3.473625
3.837083
4.241642
4.691279
5.190298
5.743359
6.355490
7.032109
7.779051
8.602583
9.509429
10.506799
11.602404
12.804489
14.121857
15.563892
17.140593
18.862595
20.741204
22.788425
25.016990
27.440391
30.072914
32.929669
36.026623
39.380639
43.009505
46.931973
51.167797
55.737767
60.663750
65.968726

p
p

1.000000
0.893114
0.799519
0.717759
0.646372
0.583987
0.529372
0.481441
0.439259
0.402023
0.369048
0.339751
0.313638
0.290289
0.269347
0.250505
0.233504
0.218120
0.204162
0.191464
0.179884
0.169297
0.159596
0.150688
0.142489
0.134929
0.127944
0.121478
0.115482
0.109912
0.104730
0.099900
0.095392
0.091179
0.087235
0.083539
0.080070
0.076811
0.073744
0.070856
0.068132
0.065561
0.063131
0.060833
0.058657
0.056595
0.054639
0.052782
0.051017
0.049339
0.047741

Table 2.6: Constant-area, supersonic flow of ideal gas with heat addition for = 1.325.

You might also like