You are on page 1of 79

1. INTRODUCTION.................................................................................................................

2. CONSERVATION OF ENERGY .......................................................................................8

2.1 The Energy Equation ..........................................................................................................8


2.2 Total (or Stagnation) Conditions ......................................................................................10
2.3 Change of State with Friction ...........................................................................................11
2.4 Mach Number ...................................................................................................................13
2.5 Total to Static Relationships for a Compressible Gas ......................................................16
2.6 X-Flow Function for a Compressible Gas ........................................................................19
2.7 Flow Areas for a Compressible Gas .................................................................................20
2.8 Examples of Application of Gas Dynamic Relationships.................................................22

3. CONSERVATION OF MOMENTUM.............................................................................25

3.1 Stream Thrust....................................................................................................................25


3.2 Expression for Stream Thrust ...........................................................................................26
3.3 Relationship of Stream Thrust to other concepts ..............................................................27

4. CONSERVATION OF MASS ...........................................................................................29

4.1 Flow-area relationship in reversible adiabatic flow..........................................................29


4.2 Duct Shape ........................................................................................................................29
5. THE NORMAL SHOCK ...................................................................................................32

5.1 Flow discontinuity in a parallel duct.................................................................................32


5.2 Normal Shock Relations for a Perfect Gas .......................................................................33
5.3 Rankine - Hugoniot Equations..........................................................................................35
5.4 Flow changes across a normal shock ................................................................................36

6. THE OBLIQUE SHOCK ...................................................................................................42

6.1 Oblique shock relations for a perfect gas..........................................................................42

7. OTHER SHOCK WAVES .................................................................................................49

7.1 Flow round a convex corner .............................................................................................49


7.2 Flow round a concave corner............................................................................................51
7.3 Detached Shocks...............................................................................................................53

1
7.4 Multiple Shocks and Expansions ......................................................................................55
7.5 Shock Reflection...............................................................................................................57

8. SHOCK—BOUNDARY LAYER INTERACTION ........................................................61

8.1 Laminar boundary layer....................................................................................................61


8.2 Turbulent boundary layer..................................................................................................62
9. FLOW IN NOZZLES .........................................................................................................64

9.1 The convergent nozzle ......................................................................................................64


9.2 The convergent-divergent (con-di) or de Laval nozzle.....................................................68
9.3 Flow in actual nozzles.......................................................................................................73
10. GAS FLOW WITH HEAT TRANSFER......................................................................76

10.1 Relationship between flow before and after heat transfer ................................................76
10.2 Physical Limitations on flow ............................................................................................77

2
PREFACE

The following notes cover the recommended minimum amount of gas dynamics needed for a
qualitative understanding of thermal propulsion systems. These notes should ideally be
augmented by your own reading; the following texts are recommended:

1. Rogers & Mayhew “Engineering Thermodynamics Work and Heat Transfer” 4th Edition,
Longman (1989)

2. Aksel & Eralp “Gas Dynamics” Prentice Hall (1994)

3. Liepmann & Roshko “Elements of Gas Dynamics” John Wiley & Sons (1957)

4. Sears (Editor) “High Speed Aerodynamics and Jet Propulsion.” Vol VI, Oxford
University Press (1955)

5. Oswatitsch “Gas Dynamics.” English Academic Press, NY (1956)

There are also obviously many internet resources available which the reader is recommended to
investigate for further reference and comprehension purposes. For more accurate numerical
work, tabulated values of many of the functions derived in these notes may be required.
Reference should therefore be made to:

6. Keenan & Kaye “Gas Tables” John Wiley NY 1945

or reference 1 above.

The reader will recognize considerable overlap with this material and that covered in the field of
“missile aerodynamics”. This is deliberate as a working knowledge of gas dynamics is of
paramount importance if the reader is to become accomplished in either the field of
aerodynamics or propulsion.

In these notes one is normally concerned with steady, uniform, one-dimensional (1-D) flow. This
may be defined as the flow in a channel or duct in which

a. The cross-sectional area is a function of the distance (X) axially along the duct (i.e. 1-D);

b. the flow is uniform over each cross-section (or, at least, over each reference section of
the duct);

c. the flow is invariant with time (i.e. steady flow).

The following summarizes the topics dealt with in the notes:

Section 1 Introduction

3
Section 2 Conservation of energy in a flowing gas and resulting concepts and deductions;

Section 3 Conservation of momentum;

Section 4 Conservation of mass in an isentropic system;

Section 5 The plane normal shock and its properties;

Section 6 The oblique shock and some of its properties;

Section 7 Miscellaneous shock formations occurring in propulsion systems;

Section 8 Shock/boundary layer interaction;

Section 9 Nozzle flow;

Section 10 Gas flow with heat.

4
NOTATION

A area, in particular, cross sectional area of duct m2

a velocity of sound m/s

c gas velocity m/s

F force or thrust N or kN

F stream thrust N or kN

G function of M & y in adiabatic flow

g gravitational acceleration m/s2

h specific enthalpy kJ/kg

ho total specific enthalpy kJ/kg

cp specific heat at constant pressure kJ/kg K

cv specific heat at constant volume kJ/kg K

m mass kg

M Mach number

p static pressure N/in2 or Pascal or bar

po total pressure N/in2 or Pascal or bar

m& mass flow kg/s

Q heat kJ

q dynamic pressure N/m2 or Pascal or bar

R gas constant kJ/kg K

S surface
s specific entropy kJ/kg K

T static temperature K

To total temperature K

u gas velocity in x direction m/s


v specific volume m3/kg
V volume m3

5
v gas velocity in y direction m/s

Ws shaft work kJ

w resultant velocity rn/s

Z height above a datum m

X,Y,Z functions of y & M

γ ratio of specific heats

δ deflection or deviation of flow through a shock

ε inclination of shock

µ Mach angle

θ flow direction from arbitrary datum


ρ density kg/rn3

Suffices

0 1 2 3 etc successive reference sections

b back value

e nozzle exit

o total

6
1. INTRODUCTION
A basic understanding of gas dynamics, or the behaviour of a gas when in motion, is of great
importance in fluid mechanics, thermodynamics and the applied fields of missile aerodynamics
and propulsion. This is because significant changes in gas properties occur when the gas is
moving relative to a fixed body. These changes become more and more prominent as the speeds
involved increase and compressibility effects become especially significant. A relevant example
of the effect of gas dynamics can be found with the high-speed flow over an aerofoil section, as
illustrated in Figure 1.1. This example is particularly important since many aeronautical or
mechanic fluid machines contain large numbers of aerofoil sections in the form of blades.

Figure 1.1 – Illustration of the effect of gas dynamics on the flow over an aerofoil

At low flow speeds the aerofoil acts as normally, generating regions of low and high pressure on
the top and bottom respectively, and thus lift. At higher flow velocities, however, the gas begins
to compress towards the front of the aerofoil and expands rapidly towards the rear, consequently
creating regions of high density gradient. Eventually, at a particular flow speed, conditions are
reached where a shock wave forms at the top of the aerofoil resulting in a severe loss of lift.
These severe density changes therefore cause the aerofoil to stall unexpectedly giving many
potential problems to the engineer (an aeronautical phenomenon often known as shock stall).
There are many other examples where significant density changes occur in gas flows, such as in
nozzles, ducts and aircraft intakes.
The following will outline the basic theory of gas dynamics and will then use the case of a
simple converging-diverging (con-di) nozzle to examine various effects of gas dynamics.

7
2. CONSERVATION OF ENERGY

2.1 The Energy Equation

In the “Introduction to Thermodynamics” notes, the First Law of Thermodynamics was applied
to a steady, one-dimensional flow in a duct with both work and heat transfer and the following
equation (14b) was obtained:

 u2
2
  u1
2

Q = Ws + m  h2 + + gZ 2  − m  h1 + + gZ1 
 2   2 

Q=0, W=0

1 2

Figure 2.1 – Schematic to illustrate isentropic flow along a duct

For the special case, as shown in Figure 2.1, in which there is no heat or work transfer between
the system and the surroundings, i.e. the flow is adiabatic and workless, and the potential energy
terms can be ignored, the above equation becomes:
2 2
u2 u
h2 + = h1 + 1 (2.1a)
2 2

It must be noted that this equation is applicable even if there are viscous stresses or any other
non-equilibrium conditions between the sections 1 and 2, as long as the reference sections 1 and
2 represent equilibrium states. If the flow is continuously in equilibrium the reference section 2
may be located at any axial position and hence:

u2
h+ = constant (2.1b)
2

Differentiating, dh + udu = 0

which is the general equation for adiabatic, zero work,equilibrium flow. If, in addition, the gas
obeys the perfect gas law pv = RT, we may write:

c p dT + udu = 0

8
If cp is constant, this equation can be integrated to give

u2
c pT + = constant (2.1c)
2

In the following sections of these notes the energy equation will frequently be written in the
form (2.1c) above, though occasionally the differential form will be adopted. If the latter form is
taken the subsequent analysis will be based on the assumption that cp is constant. This is clearly
an invalid assumption over the temperature ranges usually encountered in propulsion systems
and some justification for this treatment is necessary. It should first be noted that the treatment is
adequate for current practical calculations which can be based on one of the following
approximations:

i. A mean value of cp over the temperature range concerned can be used. This is frequently
adopted for gas turbine and supersonic ramjet engine calculations and the
thermodynamic data on the products of combustion of kerosene and air are often
presented in a way to facilitate calculations on this basis (see, for example, the data
presented in the course notes on Air-Breathing Engines).

ii. In many cases the flow is taking place at very high velocities so that the time spent in
any process, for example the flow through a rocket nozzle, is so small that the gas
properties do not respond to the changes in static conditions i.e. temperature and
pressure. In these circumstances the initial values represent a good model of the actual
flow process. Results of calculations on this basis are frequently referred to as the
“frozen equilibrium” values.

iii. Calculations may proceed on a step-by-step basis in which the thermodynamic data
appropriate to the mean conditions for each step are adopted. Provided the steps taken
are sufficiently small a very close approximation to continuous equilibrium can be
obtained. Such an approach is ideal for computer-based solutions. These calculations are
referred to as the “shifting equilibrium” or “continuous equilibrium” values, though it
should be noted that, on account of the small time available for equilibrium to be
established, it is doubtful how well this model represents actual conditions.

Provided the nature of the approximations at (i) and (ii) is kept in mind, they are adequate for
most practical purposes. Thus (i) is used where the flow velocities are low, while (ii) is adopted
at higher velocities. Should hypersonic air-breathing engines become a practical proposition,
however, a more advanced treatment will be necessary on account of the high temperatures

9
involved and the necessity to achieve very high accuracy in gross thrust calculations to retain
current accuracies in the net thrust. Since this is at present of little practical importance the
simple approach has been retained because it facilitates an analytical approach and thus enables
the physical factors to be more easily demonstrated.
2.2 Total (or Stagnation) Conditions

For an adiabatic flow system with no external work equation (2.1b) states:

specific enthalpy + specific kinetic energy = constant = total enthalpy / unit mass of gas = ho

For a gas obeying the gas law pv = RT and also having a constant specific heat cp it is possible
to define the total enthalpy in terms of a total temperature defined as follows:

specific total enthalpy = c pTo (2.2)

In this case equation (2.1c) can be written as

u2
c pT + = c pTo (2.3a)
2

u2
i.e. To = T + (2.3b)
2c p

Alternatively it may be expressed as

To = T + Tv (2.3c)

or in words

total temperature = static temperature + temperature equivalent of velocity (dynamic temperature)

Thus To = the actual temperature measured if the gas is brought to rest adiabatically (but

not necessarily isentropically)

= static temperature in a reservoir of infinite dimensions.

It will be seen later that the total temperature plays a prominent part in the development of the
theory of gas dynamics. Its importance should be self-evident even at this stage by virtue of its
close association with energy.

In conjunction with total temperature we may introduce the idea of total pressure. Thus, if the
gas is brought to rest, or is imagined brought to rest, both adiabatically and reversibly (i.e.

10
isentropically), then the resulting pressure is the total pressure, po. It will be given by (see course
notes for “Thermodynamics Fundamentals”):
γ
po  To  γ −1
= (2.4)
p  T 

Throughout these notes the phraseology “total temperature” and “total pressure” will be used.
Some text-books and papers may use the word “stagnation,” but this is not recommended. The
prefix stagnation will be reserved for the actual conditions at a stagnation point.

It should be noted carefully that the total temperature is achieved when the gas is brought to rest
adiabatically, but that the total pressure is achieved only when the gas is brought to rest
isentropically. The case when changes take place non-isentropically will be studied in sub-
section 2.3, but meanwhile it can be stated that since total and static conditions are related
isentropically, the entropy of the gas in any state is equal to the entropy at its total condition.
2.3 Change of State with Friction

When a perfect gas changes its state adiabatically (i.e. without any loss or addition of heat)
against internal friction but without doing work, then total enthalpy and hence total temperature
must remain constant.

Let an initial state be represented by

po1 total pressure, To1 total temperature, p1 static pressure

T1 static temperature, s1 specific entropy

Suppose the gas expands to a static pressure p2 in the presence of internal friction. The presence
of friction implies that the process is irreversible and that the final specific entropy s2 is greater
than its initial value s1. The initial and final states of the gas are illustrated by points 1 and 2 on
the T-s graph below (Fig 2.2) .

11
Figure 2.2 – Temperature-entropy graph for a process

The following features should be noted:

i. Total and static conditions must lie on a vertical constant-entropy line as shown, for
example, at s1 and s2 in the diagram.

ii. The difference in height between the total and static points on this vertical line is equal to
the temperature equivalent of the kinetic energy i.e. u2 / 2cp (see equation (2.3b)).

iii. If changes take place reversibly all the gas conditions would lay on this vertical line.
Thus after isentropic expansion from pressure p1 to pressure p2 the final state would be
represented by the point 1’ in the diagram

iv. If changes take place irreversibly, e.g. with friction, the new condition of the gas must lie
on another vertical line of higher entropy as shown at 2. The diagram illustrates
irreversible expansion from pressure p1 to a lower pressure p2. It is assumed to take place
along the line 1 - 2 in the diagram, but this is notional only and is shown as a dashed line
to indicate so.

v. The new static and total conditions must lie on a vertical line through 2.

vi. The new total temperature must be the same as the initial and hence lie on a horizontal
line on the diagram.

12
vii. The new total conditions are given by the intersection of the vertical line (v) and the
horizontal line (vi).

viii. It follows that, since s2 > s1, then po2 < po1

For completion, the reader could verify that, by drawing a similar diagram, conclusion (viii)
would also be reached if the expansion process discussed above, were replaced by a diffusion
process. On this basis it is concluded that, in adiabatic, zero-work flow with friction, the total
temperature will remain constant but the total pressure will decrease. The loss in total pressure
will depend on the degree of friction introduced by the process. It is customary to express the
effect of friction in terms of total pressure loss rather than entropy increase.
2.4 Mach Number

With reference to Figure 2.3, consider a fluid moving with velocity u. If there is a disturbance
upstream of the fluid, for example caused by a sudden piston movement, let the pressure wave
caused by this disturbance move through the fluid at a velocity c.

Figure 2.3 – Propagation of a pressure wave through a compressible fluid

The pressure, density and velocity upstream of the wave will be p, ρ and u while downstream of
the wave it will be p + δp, ρ + δρ and u + δu due the pressure rise increasing the local density
and imparting a force on the fluid thus accelerating it by δu. If the changes relative to the wave
position are analysed, the changes in velocity will be from (u − c) to (u − c + δu). If there is an
area ∆A across the wave, from conservation of mass:

( ρ + δρ )( u − c + δ u ) ∆A = ρ ( u − c ) ∆A
Simplifying: ( c − u ) δρ = ( ρ + δρ ) δ u
Considering the forces on the fluid and momentum (Newton’s 2nd law of motion):
Net force = Rate of change of momentum

13
( p + δ p ) ∆A − p∆A = ρ ( u − c ) ∆A ( −δ u )
Simplifying: δ p = ρ ( c − u ) δ u

 ρ + δρ  δ p
Eliminating δu leaves: ( c − u ) = 
2

 ρ  δρ

 ∂p 
(c − u )
2
For infinitesimally small changes where the flow is isentropic: = 
 ∂ρ  s

The quantity (c – u) is the speed of the pressure disturbance through the fluid relative to the
speed of the fluid. In a stationary fluid where u = 0, we call the quantity c the sonic velocity or
the speed of sound, a, such that:

 ∂p 
a2 =  
 ∂ρ  s

For a perfect gas the isentropic process is given by

p = constant ρ γ

whence a 2 = γ constant ρ γ −1

or

γp
a2 =
ρ
= γ RT (a = γ RT ) (2.5)

This is a key relationship in gas dynamics and should be memorised by the reader; it shows that,
for a given gas, the local speed of sound is proportional to the square root of the temperature.

For a compressible flow, to obtain dynamic similarity, we must scale inertial and elastic forces
such that:

[ Inertia force] ∝ ρ l 2u 2
[ Elastic force] Kl 2
where K is the bulk modulus of elasticity of a compressible fluid such that:

 ∂p   ∂p 
K = −V  = ρ 
 ∂V   ∂ρ 

u2
Hence K = ρ a 2 and we may write: C =
a2

14
This parameter is known as the Cauchy number. However, this number is equivalent to the
better known ratio Mach number defined as:

u
M= (2.6)
a

Hence the Mach number is simply a non-dimensional ratio of local velocity (u) to the local
speed of sound (a) and ensures dynamic similarly in a compressible flow. Thus it is one of the
most important parameters in compressible flow theory and also allows a sharp division in flow
properties to be clearly defined as follows:

If M<1 the flow is said to be subsonic;

when M=1 the flow is said to be sonic;

and if M>1 the flow is said to be supersonic.

It should be noted that the Mach number M may vary from point to point in a fluid both because
u varies and because a itself varies according to equation (2.5).

Example 1

A missile is travelling at a velocity of 500 m/s at an altitude where the pressure and
density are 70 kPa and 0.909 kg/m3 respectively. Meanwhile, the wind is blowing in the
opposite direction at a velocity of 35 m/s. Determine: (i) the wind Mach number, (ii) the
missile Mach number if there were no wind, and (iii) the missile Mach number relative
to the moving air.

Example 1 - Solution

Using the equation of state: T = p / (ρ R) = 70000 / (287 x 0.909) = 268.2 K

Using equation (2.5): a = √(γ R T) = √ (1.4 x 287 x 268.2) = 328.3 m/s

Using equation (2.6): (i) wind M = 35 / 328.3 = 0.1066

(ii) missile M (no wind) = 500 / 328.3 = 1.523

(iii) missile M (with wind) = 535 / 328.3 = 1.63

The major significance of a supersonic flow is that, when the gas is moving at greater than the
speed of sound, pressure waves cannot be propagated upstream to influence the flow. This will
also be appreciated further once the reader has developed his knowledge of missile
aerodynamics. In contrast, in a subsonic flow, the pressure waves can be propagated both

15
upstream and downstream. Figures 2.4 and 2.5 illustrate this effect.

u
u < a (M < 1)

u
u > a (M > 1)

Figure 2.4 – Effect of Mach number on the flow around a solid object

u < a (M < 1) a

u > a (M > 1) a

Figure 2.5 – Effect of Mach number on the flow inside a nozzle

2.5 Total to Static Relationships for a Compressible Gas

We are now in a position to establish analytical relationships between the total and static
conditions at any point in a flowing compressible fluid.

For a perfect gas, equation (2.3) gives:

16
u2  γ −1   γ −1  u
2
u2 T u2
c pTo = c pT + or o = 1 + = 1+   = 1 +   2
2 T 2c pT 2T  γ R   2 a

i.e.

To  γ −1  2
= 1+  M (2.7)
T  2 

Additionally, making use of equation (6) we can write


γ γ
po  To  γ −1
 γ −1 2  γ −1
=  = 1 + M  (2.8)
p T   2 

Finally, from the isentropic relationship

1 1
γ
ρo  po   γ −1 2  γ −1
= = 1 + M  (2.9)
ρ  p   2 

Equations (2.7), (2.8) and (2.9) provide the relationships between the total and static
temperature, the total and static pressure, and the total and static density, all in terms of γ and M,
at any point in a flowing fluid. They should all be committed to memory by the reader as they
will be used in the majority of missile propulsion (and indeed aerodynamics) calculations later
on. For convenience, equations (2.7) and (2.8) may be rewritten as:

To po
= Z (γ , M ) and = Y (γ , M )
T p

These are known as the Z and Y functions respectively. Curves of To / T and po / p are plotted
against Mach No for a practical range of γ in Figure 2.6 and Figure 2.7.

17
Figure 2.6 – Total to static temperature ratio To/T versus Mach No M (Z-function)

Figure 2.7 - Total to static pressure ratio po/p versus Mach No M (Y function)

18
2.6 X-Flow Function for a Compressible Gas

The mass flow at any section is given by m& = ρ Au (Continuity equation)

T  M2
where u = Ma = M γ RT = M γ RTo   = γ RTo
 To   γ −1  2
1+  M
 2 

ρo po 1
and ρ= 1
= × 1
  γ −1  2  γ −1 RTo   γ − 1  γ −1
2
1 +  2  M  1 +   M 
      2  

po 1 M2
∴ m& = × × A × γ RT
 γ −1  2
1 o
RTo   γ − 1  γ −1
2 1+ 
1 +   M M
   2 
  2  

This can be rearranged to give:

m& RTo M γ
∴ = (γ +1)
= X (γ , M ) (2.10)
Apo
  γ −1   2(γ −1)

1 +  2  M 
2

   

This is commonly known as the X-flow function and is widely used in gas dynamics
applications, as will be appreciated later. Values of X (γ, M) have been plotted against M for the
practical range of γ in Figure 2.8.

It should be noted that X(γ, M) increases with increasing M from 0 to Mach l. It reaches a
maximum at Mach l irrespective of the value of γ; this can be verified by differentiating equation
(2.10) with respect to M and equating to zero. The only maximum occurs at Mach l, and above
that value X (γ, M) decreases steadily to value zero at M∞ (provided that the assumptions on
which equation (13) has been deduced hold over this range). The reason for this shape is the
rapid decrease of density with increasing Mach number (see equation (2.9)) necessitating an
increasing flow area with increasing Mach number. Above Mach l this effect becomes greater
than the effect of increasing velocity u, which demands a decreasing flow area with increasing
Mach number. It follows from the shape of these curves that:

i. for any value of X there are in general two values of M, one for subsonic flow and one
for supersonic flow, satisfying the flow/area relationship; to establish which of these
flow conditions prevails it is necessary to examine either the total-to-static pressure ratio

19
or the total-to-static temperature ratio;

ii. a unique value of Mach number occurs at the maximum value of X. At this point the
Mach number is equal to unity and for given values of the mass flow m& and gas
conditions p0 and To the flow area is a minimum.

The consequences of these factors on duct design will be examined in detail in Section 4.

Figure 2.8 – Plot of the X flow function versus Mach No M

2.7 Flow Areas for a Compressible Gas

It is often convenient to express the flow area in terms of the flow area at Ml which will be
denoted by At (The reason for this nomenclature will be discussed in section 4.2.

At Mach number M:

m& RTo M γ
= (γ +1)
, (2.10)
Apo
  γ −1   2(γ −1)

1 +  2  M 
2

   

At Mach number 1:

20
m& RTo γ
= (γ +1)
(2.10a)
At po
 γ − 1   (γ −1)
2

1 +  2  
  

Hence it may be shown that:

(γ +1)
1  2  ( γ − 1) M   2( γ −1)
2
A
=  1 +  (2.11)
At M  ( γ + 1)  2  

The ratio A/At has been plotted against M for a range of γ in Figure 2.9. Like X this is a two-
valued function. It has a minimum at Mach 1. This is an extremely useful relationship, allowing
the flow Mach number to be calculated at any point along a duct of known geometry, provided
the velocity is known at a given point.

Figure 2.9 – Graph to show area ratio A / At versus Mach No M

21
2.8 Examples of Application of Gas Dynamic Relationships

Given below are two examples, illustrating how the relationships derived in sub-sections 2.5, 2.6
and 2.7 may be applied to practical gas dynamic problems, in particular to thermal propulsion
systems.

Example 2

Find the air (γ = 1.4) static temperature at inlet to the combustion chamber of a ramjet
flying at Mach 2 at an altitude of 15000 m (ambient static temperature = 216.65 K) if the
Mach number at the combustion chamber inlet is 0.25.

Figure 2.10 – Example of ramjet conditions

Example 2 - Solution

At M = 2 for air, the ratio To / T is obtained from Figure 2.6:

∴Toa = 1.8 x 216.65 = 390 K

  1.4 − 1  2 
Or directly from equation (2.7): T0 a = Ta 1 +   2  = 216.65x1.8 = 390 K
  2  

Since the flow from the atmosphere to the chamber entry is both adiabatic and workless,
the enthalpy and therefore the total temperature remain constant, i.e. To2 = 390K

At M = 0.25 for air, using equation (2.7), To / T = 1 + (γ-1)/2 (0.25)2 = 1.0125

∴Temperature of the air at 2 = 390 / 1.0125 = 385K

Example 3

A liquid rocket engine has a total propellant mass flow of 5 kg/s. The temperature and
pressure after combustion are 3000 K and 20 bar respectively. The products of

22
combustion have a gas constant R = 0.360 kJ/kg K and γ = 1.25. What must be the
nozzle exit area for maximum thrust? What is the thrust then obtained?

Figure 2.11 – Example of rocket motor conditions

Example 3 - Solution

Note that the combustion chamber velocity is necessarily low therefore it can be
regarded as an infinite reservoir in which the velocity is zero. In that case

total pressure after combustion = 20 bar

total temperature after combustion = 3000 K

Neglecting frictional losses in the nozzle and heat losses through the nozzle wall (in
practice very small)

total pressure at nozzle exit = 20 bar

total temperature at nozzle exit = 3000 K

For maximum thrust the static pressure at the nozzle exit should be equal to ambient
pressure (see Rocket Propulsion Notes).

∴static pressure at nozzle exit = 1 bar

∴ po / p at nozzle exit = 20

∴ Mach No at nozzle exit = 2.55

To find the flow area (from Figure 2.8 for γ = 1.25),

m& RTo
= X ( M , γ ) = 0.2 at nozzle exit
Ae po

(from Figure 2.8 for M = 2.55 and γ = 1.25, or by using equation (2.10) directly)

23
m& RTo 5 360 × 3000
Ae = = = 13 x 10-3 m2
po X ( M , γ ) 20 × 105 × 0.2

To obtain the thrust, the velocity at the nozzle exit is required. This is given by M × a,
where a is the velocity of sound given by:

a = γ RT , T being the local static temperature.

At M = 2.55, To / T = 1.82 (from Figure 2.6 for γ = 1.25 or using equation (2.7))

∴ T = 3000/1.82 = 1648 K, and a = 1.25 × 360 × 1648 = 860 m/s

The velocity at nozzle exit is then equal to M × a = 2.55 x 860 = 2200 m/s

Since the exit and ambient pressures are equal:

Thrust = mu
& e = 5 x 2200 = 11 kN

24
3. CONSERVATION OF MOMENTUM
3.1 Stream Thrust

Consider a gas flowing in a duct of arbitrary cross section in the presence of friction. The forces
acting on the fluid between the sections 1 and 2 are shown in the Figure 3.1 below:

Figure 3.1 – Nozzle modeling parameters

Applying Newton’s second law where f is the frictional force per unit length leads to,

Rate of change of momentum = applied force


2 2
∴ m& ( u2 − u1 ) = p1 A1 + ∫ pdA − ∫ fdl − p2 A2
1 1

2 2

The terms  ∫ pdA − ∫ fdl  represent the total axial force exerted either
1 1 

i. by the bounding surface on the fluid from 1 → 2 or from Newton’s third law

ii. by the fluid on the bounding surface from 2 → 1


2 2
Thus the axial force F, is given by F = ∫ pdA − ∫ fdl
1 1

25
which then becomes F = ( mu
& 2 + p2 A2 ) − ( mu
& 1 + p1 A1 )

This may be written as

F = F%2 − F%1 (3.1a)

In which F% is used as a general quantity defined as

F% = mu
& + pA (3.1b)

The quantity F% is called the stream thrust. It is of considerable practical importance, since the
change in stream thrust between two sections gives the total axial force exerted (in a direction
contrary to the gas flow) by the gas on the boundary, whatever the changes of state and flow
within the boundary may be.
3.2 Expression for Stream Thrust

In addition to its practical importance the stream thrust is a convenient quantity to use in as
much as it may be expressed in terms of γ and M. Thus

γp 2
& + pA = ρ Au 2 + pA = ρ Aa 2 M 2 + pA = ρ A
F% = mu M + pA = pA ( γ M 2 + 1)
ρ

Expressing the static pressure in terms of po , γ, M

γ M 2 +1
F% = po A γ (γ −1)
 γ −1 2 
1 + 2 M 

This is often written as

F% = po AG (3.2)

where

γ M 2 +1
G= γ (γ −1)
(3.2a)
 γ −1 2 
1 + 2 M 

G is plotted in Figure 3.2 for γ = 1.4 and 1.25 against Mach number. An enlarged version of the
part of the curve lying between M = 0.1 and M = 0.8 is given in the insert covering combustion
chamber applications in particular.

26
Figure 3.2 – G function versus Mach No for various γ

3.3 Relationship of Stream Thrust to other concepts

In the absence of friction the general momentum expression can be written

m& ( u2 − u1 ) = p1 A1 − p2 A2 + ∫ pdA

& = − d ( pA) + pdA


or in differential form as mdu

i.e. ρ Audu = − pdA − Adp + pdA

− dp
or udu = (3.3)
ρ

which is Euler’s equation for steady one-dimensional flow.

u2 dp
Integrating gives +∫ = constant , which is the compressible form of Bernoulli’s equation.
2 ρ

If ρ is constant we get the familiar

u2 p
+ = constant (3.4)
2 ρ

It is thus seen that the stream thrust is an alternative form of either Euler’s or Bernoulli’s

27
equations. Both will be covered and used in extremely more depth in the accompanying
aerodynamics notes and lectures. Since we are essentially interested in thrust for propulsion
purposes, the stream thrust is a much more convenient form to use for this application. This
derivation also emphasizes that Bernoulli’s equation is fundamentally based upon a momentum
relationship and is not really an energy equation. The energy equation is that derived in Section
2.1, though it should be noted that in incompressible isentropic zero-work flow the energy
equation will reduce to the familiar Bernoulli form anyway.

28
4. CONSERVATION OF MASS
4.1 Flow-area relationship in reversible adiabatic flow

Consider the isentropic flow of a gas in a duct of arbitrarily varying cross-section. Conservation
of mass is represented by the following well-known equation:

m& = ρ Au = constant (4.1)

Writing equation (4.1) in logarithmic form and differentiating we obtain:

dρ dA du
+ + =0
ρ A u

Which can be re-written as: udu + c p dT = 0

Since the flow is also reversible since pvγ = const = p ρ −γ and pv = RT = p ρ −1 :

p dρ
RT = = const ρ γ −1 ∴ RdT = ( γ − 1) const × ρ γ − 2 d ρ = ( γ − 1) RT
ρ ρ

γ dρ dρ
∴ c p dT = R (γ − 1) T and c p dT = a 2
γ −1 ρ ρ

a 2d ρ dρ udu du
Substituting for cp dT, udu = − or =− = −M 2
ρ ρ a 2
u

Back-substituting we obtain

dA du
= − (1 − M 2 ) (4.2)
A u

This is an extremely useful relationship, as will be demonstrated later, and must be satisfied in
isentropic flow in any duct.
4.2 Duct Shape

It follows from equation (4.2) that

i. If the flow is subsonic (i.e. 0 < M < 1) then the duct cross-sectional area has to decrease
if the flow velocity is to increase and vice versa (see Figure 4.1).

ii. If the flow is supersonic (i.e. M > 1) then the duct cross-sectional area has to increase if
the flow velocity is to increase and vice versa (see Figure 4.2).

iii. The conditions for the flow to be sonic (i.e. M = 1) can be seen from a duct which is

29
required to accelerate the flow from a low subsonic Mach No to a supersonic Mach No.
In the subsonic region the duct converges. In the supersonic region the duct diverges (see
Figure 4.3. Hence, in the region where M = 1, there will be a throat in accordance with
equation (4.2), which gives dA = 0 for M = l.

Figure 4.1 – Subsonic flow (i.e. 0 < M < 1) then decrease in duct cross-sectional area gives
increase in flow velocity and vice versa

Figure 4.2 – Supersonic flow (i.e. M > 1) then increase in duct cross-sectional area gives
increase in flow velocity and vice versa

30
Figure 4.3 – A convergent divergent duct is required to accelerate the flow from a low subsonic
Mach No to a supersonic Mach No

It should be noted from equation (4.2) that if M = 1 there must be a throat since dA/A = 0.
Additionally, for M to be equal to unity, dA/A must be zero if du/u is to remain finite. Thus M =
1 can only be attained at a throat. It follows that a continuous flow passage which is to accelerate
a subsonic flow adiabatically and reversibly to a supersonic flow must first converge and then
diverge. At the throat of such a convergent-divergent duct (or nozzle) M = l. Likewise, a
continuous flow passage used to decelerate a supersonic flow adiabatically and reversibly to a
subsonic flow must first converge and then diverge. Again, at such a throat, M = l.

For either of the flow regimes listed in (iii) to be established the total/static pressure ratio and the
flow function at the throat must be satisfied:

γ γ −1
 po   γ +1
Using equation (2.8):   = Y (γ ,1) =   (4.3)
 p t  2 

 m& RTo  γ
Using equation (2.10):   = X (γ ,1) = (γ +1)
(4.4)
 Apo t  γ + 1  2(γ −1)
 
 2 

It should be noted that, if equations (4.3) and (4.4) are not satisfied, sonic conditions will not be
set up in the throat. In these circumstances equation (4.2) merely states that at such a throat du is
zero and the flow velocity has reached a maximum or minimum, i.e. if subsonic will have
accelerated to the throat and will be diffusing downstream. The reverse would hold in supersonic
flow. Thus the flow in a geometric throat in which equations (4.3) and (4.4) are not satisfied
may be either subsonic or supersonic.

31
5. THE NORMAL SHOCK
5.1 Flow discontinuity in a parallel duct
Consider in the first instance the adiabatic zero-work flow in a parallel frictionless duct,
concentrating on the flow between two arbitrary sections 1 and 2 as in Figure 5.1 below.

Figure 5.1 – Schematic of shock control volume

Since the area at sections 1 and 2 is the same, the equations for conservation of mass,
momentum and energy become:

Mass : ρ1u1 = ρ2u2

Momentum : p1 + ρ1u1 = p2 + ρ 2u2


2 2

2 2
u u
Energy : c pT1 + 1 = c pT2 + 2
2 2

γ γ  p1 p2 
Now c p ( T1 − T2 ) = R (T1 − T2 ) = −
γ −1 γ − 1  ρ1 ρ 2 

2 2 γ  p1 p2 
And back-substituting into the momentum equation gives: u2 − u1 = −
γ − 1  ρ1 ρ2 

Eliminating p2 and ρ2 gives:

2 2γ  p1   2γ  p1 γ − 1 2
u2 −  + 1 u1u2 +   +  u1 = 0
γ + 1  ρ1u12
  γ + 1  ρ1u1 γ + 1 
2

This is a quadratic equation for which there are two solutions.

32
u2 = u1

 2γ  p1 γ − 1
and u2 =   +  u1 (5.1)
  γ + 1  ρ u
1 1
2
γ + 1 

The first solution is the trivial solution in which the flow remains unchanged from section 1 to
section 2. The second solution is real and represents an alternative flow regime. It may be noted
that there are no restrictions on the locations of sections 1 and 2. 2 may therefore be allowed to
approach 1, in which case there will be a discontinuity between the flow before and after the
plane 1-2. Such a discontinuity is known as a shock, and, since it is normal to the flow direction,
usually called a normal shock. The circumstances in which a jump can occur, i.e. a normal
shock can be formed, will be discussed in subsequent sections. It is first necessary to establish
the flow conditions before and after a shock. The velocity ratio has already been deduced and is
given by equation (23) which may be expressed in any of the following forms:

u2  2γ  p1 γ − 1
=   +  (5.2a)
u1  γ + 1  ρ1u1 γ + 1
2

u 2  2  1 γ −1
i.e. =   2+  (5.2b)
u1  γ + 1  M 1 γ + 1 

u2  2 + (γ − 1) M 1 
2

or =  (5.2c)
u1  (γ + 1) M 12 

Alternatively:

1   γ −1  2 
1+ M
u2 2
 2 + ( γ − 1) M 1  2 2 2   2  1  2
2

u1u2 = × u1 =   M 1 a1 = a1 (5.3)
 ( γ + 1) M 1  1   γ − 1 
2
u1
1+  
2   2  

5.2 Normal Shock Relations for a Perfect Gas

The relationships between the other gas conditions before and after the normal shock can now be
obtained and expressed in terms of γ and M as follows:

a. The static pressure ratio can be deduced from the momentum equation by rewriting as

 ρ 2 u2 2 
p1 − p2 = ρ1u1 − ρ 2u2 = ρ1u1 1 −
2 2 2
2 
 ρ1u1 

33
2  u  p2 − p1 ρ1u1  u2 
2

Substituting from the mass equation: p2 − p1 = p1u1 1 − 2  or = 1 − 


 u1  p1 p1  u1 

Substituting for ρ1/p1 = γ/a12 and for u2 / u1 from equation (5.2c)

p2 − p1  2 + (γ − 1) M 12 
= γ M 1 1 −
2

p1  (γ + 1) M12 
2γ 2γ γ −1
or
p2
p1
= 1+
γ +1
2
(
M1 −1 =
γ +1
) 2
M1 −
γ +1
(5.4a)

Another useful form in terms of both M1 and M2 may be derived:

p2 1 + γ M 1
2

or = (5.4b)
p1 1 + γ M 2 2

b. The density ratio follows immediately from the continuity equation since

ρ 2 u1
= =
( γ + 1) M 12
(5.5)
ρ1 u2 2 + ( γ − 1) M 12

c. The static temperature ratio can now be determined from the perfect gas law

T2 p2 ρ1
=
T1 ρ 2 p1

Substituting for p2 / p1 from equation , ρ1 / ρ2 from equation (5.5):

T2  2γ M 1 − (γ − 1)   2 + (γ − 1) M 1 
2 2

= (5.6a)
(γ + 1) M12
2
T1

Another useful form in terms of both M1 and M2 may be derived:

1
1 + (γ − 1) M 1
2
2 + (γ − 1) M 1
2
T2 2
or = = (5.6b)
T1 1 + 1 (γ − 1) M 2 2 + (γ − 1) M 2 2
2
2

d. The Mach number after the shock is obtained from the velocity ratio, and the static
temperature ratio which controls the velocities of sound before and after the shock. Thus
2 2
u1  2 + ( γ − 1) M 1   2 + (γ − 1) M 12 
2 2 2 2
2 u2 2 2a
M2 = 2 = 2  or M2 = M 12  
a2  (γ + 1) M 1   (γ + 1) M 1 
2 1 2
a1 a2

34
2 2  2γ M 12 − (γ − 1)   2 + (γ − 1) M 12 
Since 2 = , equation (5.6) can be used to give: 2 =   
a2 T2 a2
(γ + 1) M 1
2 2
a1 T1 a1

and back-substituting gives:

2 + (γ − 1) M 1
2
2
M2 = (5.7)
2γ M 1 − (γ − 1)
2

e. The gas total conditions after the normal shock are obtained from the static conditions given
by a, b and c above plus the relationships of section 1.5, remembering that the total temperature
will remain constant. Thus

γ (γ −1) γ (γ −1)
T  T 
T02 = T01 , p01 = p1  o1  , p02 = p2  o 2 
 T1   T2 

γ (γ −1)
p02 p2  T1 
Hence: =  
p01 p1  T2 

Substituting for p1 from equation (5.4a) and T1 from equation (5.6a):

γ (γ −1)
( γ − 1)   ( γ + 1) M 12 
2
po 2  2γ 2 
= M1 −  2 
(5.8a)
po1  ( γ + 1) (γ + 1)    2γ M1 − (γ − 1) 2 + (γ − 1) M 1  
2

Or, alternatively, it may be shown that:

γ ( γ −1) 1 ( γ −1)

po 2  ( γ + 1) M 1 
2  ( γ + 1) 
 
=    (5.8b)
po1  2 + ( γ + 1) M 12    2γ M 1 − ( γ − 1)  
2

5.3 Rankine - Hugoniot Equations

Proceeding as in section 5.1, but eliminating u1 and u2 from the equations to obtain a relationship
between the pressure and density, the Rankine-Hugoniot Equations are obtained. They may
alternatively be derived by eliminating M1 from equations (5.4) and (5.5). Either approach gives

 γ + 1  ρ2
  −1
p2  γ − 1  ρ1
= (5.9)
p1  γ + 1  ρ 2
 γ −1  − ρ
  1

35
 γ + 1  p2
1+ 
ρ2  γ − 1  p1
= (5.10)
ρ1  γ + 1  p2
 γ −1  + p
  1

The pressure-density relationship given by the Rankine-Hugoniot Equations is presented in


Figure 5.2 and compared with the isentropic relationship below:

Figure 5.2 – The Rankine-Hugoniot Relationship

It will readily be seen from this graph of the Rankine-Hugoniot relationship that the flow across
a normal shock is not isentropic, i.e. it is not reversible. It follows that any freedom for the flow
to jump across the shock and then back to its original state cannot be allowed. For a closer
examination of this feature, the second law of thermodynamics must be revisited, as in the
following sub-section.
5.4 Flow changes across a normal shock

The change of entropy between two states is given in the “Thermodynamics” notes by:

T2 p γ p ρ  p
s2 − s1 = c p ln − R ln 2 = R ln  2 1  − R ln 2
T1 p1 γ −1  ρ 2 p1  p1

36
s2 − s1  p 1 (γ −1)  ρ γ (γ −1) 
i.e. = ln  2   
1
 (5.11)
R  1 
p ρ
 2 

This general relationship may be applied to the particular case of the discontinuity under
discussion by substituting for p2 / p1 from equation (5.4a) and p1/p2 from equation (5.5) to give:

 1 ( γ −1) γ (γ −1) 
s2 − s1  2γ γ −1   2 + ( γ − 1) M 12 

= ln 1 +
2
M1 −   
R  γ + 1 γ + 1   ( γ + 1) M
2
 
 1

where s1 and s2 are the entropies before and after the discontinuity respectively. This expression
can be rewritten in terms of (M12 - 1) as follows:

γ ( γ −1)
 1 ( γ −1)
 1 
γ (γ −1)
 ( γ + 1) − ( γ − 1) − ( γ − 1) M 12  
s2 − s1  2γ  
R
= ln  1 +
 γ + 1
( 2
M1 − 1 

)  2 
 ( γ + 1)



  M1  

γ ( γ −1)
 2γ 
1 ( γ −1)
( γ −1)  ( γ − 1) 2  
 −γ

= ln  1 + (
2
M1 − 1  ) 1 + M 12 − 1 
  ( ) 1 +
( )
M1  
  γ + 1   γ + 1  

ε2 ε3 ε4
If (M12 - 1) is small this can be expanded using ln (1 + ε ) = ε − + − + ...
2 3 4

(M )
2 3
s2 − s1 2γ 1 −1
∴ = + ... (5.12a)
(γ + 1)
2
R 3

An alternative form is:

 2γ M 2 − ( γ − 1)  2 + ( γ − 1) M 2  γ 
 
∴ s2 − s1 = cv ln  1

1
  (5.12b)
 ( γ + 1)  ( γ + 1) M1   2

37
Figure 5.3 – Theoretical Mach No relationships before and after a normal shock

For (s2 – s1) to be greater than zero, (M2 - 1) must be positive which gives M12>1 or: M>1. Thus
the second law of thermodynamics demands that the flow before the discontinuity, or normal
shock, be supersonic. It follows that only those portions of the relationships established in
sections 5.1, 5.2 and 5.3 may be applied to a supersonic inlet velocity. This condition imposes a
number of limitations on the flow that can be established. For example, an unrestricted use of
equation (5.7) would give the Mach numbers after the shock shown in Figure 5.3.

However, if the Mach number before the shock is restricted to values greater than unity it is seen
that the Mach number after the shock must be restricted to values less than unity, i.e. the flow
after a normal shock must be subsonic.

The ratios of the major significant fluid properties before and after a normal shock have been
deduced from equations (5.4), (5.7) and (5.8) for M1 > 1 and γ = 1.4 (i.e. the curves apply to air)
and are plotted on Figures 5.4 – 5.7. It will be noted that the static pressure ratio is always
greater than unity, so that the process is a compression, but on account of the non-reversibility
the total pressure ratio is less than unity. The properties of the normal shock may therefore be
summarized as follows:

l. A normal shock can only occur when the Mach number before the shock is greater than
unity.

2. The Mach number after the shock will be less than unity.

38
3. The shock is a non-isentropic process.

4. The static pressure rises across the shock resulting in a compression.

5. As M → 1.0 the process approaches isentropic and the static pressure ratio approaches
unity. So too does the total pressure ratio.

By virtue of the cube relationship of equation (5.12), it may be noted that the increase of entropy
across the shock becomes rapidly larger at the higher Mach numbers. Consequently the total
pressure loss across the shock increases rapidly with Mach number. As an illustration, the total
pressure loss across the shock is only some 7% at a Mach number of 1.5 but increases to

Figure 5.4 - Mach No relationships before and after a normal shock

39
Figure 5.5 – Total pressure before and after a normal shock

Figure 5.6 – Static temperature before and after a normal shock

Figure 5.7 – Static pressure before and after a normal shock

nearly 30% at a Mach number of 2 and has become nearly 70% at a Mach number of 3. Since
pressure losses much in excess of 10% cannot be accepted in an engine, the normal shock is
generally to be avoided even though it offers a remarkably compact method of reducing the
speed of a supersonic streamtube and achieving high compressive static pressure ratios (the
static pressure ratio is 4.5 at M = 2 which is adequate for a simple jet engine, rising to 10 at M =
3). Provided the Mach number at which it occurs can be restricted to something in the region of

40
1.5 or less, the normal shock is, however, then a practically effective compression process. This
theory is relevant when it comes to considering the different types of intakes suitable for air-
breather propulsion units at supersonic speeds.

Example 4

An air-stream with a velocity of 500 m/s, a static pressure of 70 kPa and a static
temperature of 300 K undergoes a normal shock. Determine: (a) the Mach number and
velocity after the normal shock, (b) the static conditions after the normal shock, (c) the
total (i.e. stagnation) conditions after the normal shock, and (d) the entropy change
across the normal shock.

Example 4 – Solution
(a) Using equation (5.7) or Fig 5.4: M22 = (2 + (1.4 – 1) M12) / (2 x 1.4 x M12 – (1.4 – 1))
∴ M2 = 0.723
M1 = 500 / √(1.4 x 287 x 300) =1.440
Using equation (5.5): u2/u1 = ((1.4 + 1) M12) / (2 + (1.4 – 1) M12)
∴ V2 = 284.3 m/s
(b) Using equation (5.4a) or (5.4b): p2 / p1 = 2.253, ∴ p2 = 157.7 kPa
Using equation (5.6a) or (5.6b), T2 / T1 = 1.281, ∴ T2 = 384.3 K
(c) Total pressure from equation (2.8): p1 / po1 = 0.2969, ∴ po1 = 235.8 kPa
Using equation (2.7): T1 / To1 = 0.7069, ∴ To1 = 424.4 K
Using equation (5.8): po2 / po1 = 0.9476, ∴ po2 = 223.4 kPa
Since the flow is adiabatic, To2 = To1 = 424.4 K
(d) Entropy change from equation (5.12): s2 – s1 = 57.55 J/kgK

41
6. THE OBLIQUE SHOCK
6.1 Oblique shock relations for a perfect gas

A more common and practical case than a normal shock is when a supersonic flow is forced to
change direction, by passing over a sharp concave corner, thus forming a shock inclined to the
initial flow direction and known as an oblique shock wave. The oblique shock can be treated
directly from the equations of motion. Equations may also be derived directly from the normal
shock by superimposing a velocity component parallel to the shock, as shown in Figure 6.1.

Figure 6.1 – Oblique shock geometry

The derivation and supporting theory behind all of the fundamental equations associated with
oblique shocks are deemed to be beyond the scope of these notes, which are, after all, intended
mainly for “Propulsion” purposes, rather than “Aerodynamics”. Good accounts may be found in
many specialist Gas Dynamics text books, e.g. “Gas Dynamics” by Zucrow and Hoffman,
Wiley, 1976.

These notes will merely present the most useful equations and some examples. All are based on
an assumption of the flow passing over a frictionless solid wall. The theory is based on
decomposing the flow into components which are normal and tangential to the oblique wave so
that, effectively, the shock may be considered to be a normal type which affects the normal flow
component only.

For the static pressure ratio, we obtain:

p2 2γ  γ −1 
M 1 sin 2 ε − 
2
=  (6.1a)
p1 γ + 1  γ +1 

42
For the density ratio:

ρ 2 tan ε
= =
( γ + 1) M 12 sin 2 ε
(6.1b)
ρ1 tan β 2 + ( γ − 1) M 12 sin 2 ε

The velocity ratio is given by:

u2 sin ε
=
 2
+
(γ − 1) 
  (6.1c)
u1 sin β  ( γ + 1) M 1 sin ε ( γ + 1) 
2 2

The Mach number after the shock wave is obtained from:

tan ε  2  
= 
1 ( γ − 1) 
  2 2 +  (6.1d)
tan β  ( γ + 1)   M 2 sin β 2 

The oblique shock wave angle may be calculated from:

1  ( γ + 1)   M 12 
=   2 2 − 1 tan ε (6.1e)
tan δ  2   M 1 sin ε 

When ε = π/2 the above equations transform to those appropriate for a normal shock, as
presented earlier in Chapter 5.
Once the Mach numbers across the oblique shock are known, the total pressure ratio can be
calculated from the static pressure ratio. Several useful graphical versions of these equations
may be prepared, as shown in Figs 6.2 to 6.4, all for air (γ = 1.4).

It may be observed from the charts that, for any given initial Mach number M1 and flow
deviation δ, there are two values of the exit Mach number and static and total pressure ratio. The
one with the smaller inclination and smaller static pressure ratio is usually referred to as the
weak shock and the one with the higher inclination and static pressure ratio as the strong shock.
The boundary conditions will dictate which type will occur, though the weak shock is generally
more commonly found in practice.

The Mach number at exit from the strong shock is always less than 1, i.e. the flow is subsonic.
The Mach number at exit from the weak shock is generally greater than 1, though it can be less
than 1 in some cases. At any given value of the entry Mach number (M1), the strong and weak
shocks approach each other as the deviation (δ) is increased, and become coincident and
identical when the deviation reaches its maximum value (denoted by δmax). There is no simple
solution for the flow pattern satisfying the conditions for which these charts have been prepared

43
if a deviation greater than the maximum is imposed on the flow. At the other end of the range,
represented by the limiting case of zero deflection, the weak shock becomes a Mach line
inclined at the Mach angle (see) and the strong shock becomes the normal shock.

Figure 6.2 –Oblique shock wave angle (ε) as function of M1 and deflection (δ)

44
Figure 6.3 – Downstream Mach number (M2) for oblique shock as
function of M1 and deflection (δ)

45
Figure 6.4a – Static pressure ratio across oblique shock as

function of M1 and deflection (δ)

It will be appreciated from this treatment of the oblique shock as a normal shock with transverse
flow that the flow at exit perpendicular to the shock will be subsonic and that there will be an
increase of entropy corresponding to this perpendicular flow. In all cases there is therefore a loss
of total pressure. The values may be obtained from the graphs or by using the appropriate
equations. It will be seen from these curves that the loss is greatest through a normal shock, less
through a strong shock and still less through a weak shocks, becoming zero through a Mach line

46
which is the limiting isentropic case.

The entropy increase across the oblique shock may be calculated from:

  p   ρ γ 
∆s = cv ln  2   1   (6.2)
 p1   ρ 2  

Figure 6.5b – Total pressure ratio across oblique shock as

function of M1 and shock angle (ε)

The treatment given in this section is essentially two-dimensional, i.e. applies to plane oblique

47
shocks. The treatment would not apply generally in three dimensions, and in particular would
not apply to conical shocks which are often generated in axisymmetric flow. Treatment of this
case is outside the scope of these notes.

Example 5

A uniform supersonic air flow traveling at a Mach number of 3.0 passes over a concave
corner, as shown below. An oblique shock wave, which makes an angle of 30o with the
initial flow direction, is attached to the corner under the given conditions. If the pressure
and temperature of the uniform flow are 25 kPa and -50oC respectively, determine the
pressure and temperature behind the wave, the downstream Mach number and the
deflection angle.

Example 5 – Solution
The normal component of the upstream Mach number is 3.0 x sin 30o = 1.50.
From equation (5.4): p2 / p1 = 2.458, ∴p2 = 2.458 x 25 = 61.45 kPa
From equation (5.6): T2 / T1 = 1.320, ∴T2 = 1.32 x 223.15 = 294.56 K
From equation (5.8): M2n = 0.7011
For an adiabatic process, To1 = To2, then:
From equation (2.7), for M1 = 3.0: To1/T1 = 2.8, ∴To1 = To2 = 624.5 K
∴ T2/To2 = 294.56 / 624.5 = 0.4714 and, from equation (2.7): corresponding M2 = 2.367
From Fig 6.1, sin(ε − δ ) = u 2 n / u 2 = M 2 n / M 2 = 0.7011/2.367

∴ε-δ = 17.21o and δ = 30o – 17.21o = 12.79o

48
7. OTHER SHOCK WAVES
7.1 Flow round a convex corner

When a supersonic flow is caused to change its direction, the cause of the change becomes a
source of disturbance. Consider the flow round a convex corner:

Figure 7.1 – Flow geometry around a corner

At reference plane (1) the flow is parallel to surface and at (2) parallel to S2. Because area A2 is
effectively greater than A1 the flow accelerates. Hence

u2 > u1 
 i.e. an expansion
p2 < p1 

a. Consider first an infinitely small expansion originating at 0.

Figure 7.2 – Single expansion geometry

This may be treated as a reversed shock approaching zero intensity. Downstream of the
disturbance 0-L the Mach number will be greater, and hence the Mach angle µ less than before
the disturbance. The flow will be turned towards the disturbance.

49
b. Now consider a series of small expansions originating at 01 02 03 etc

Figure 7.3 – Geometry for a series of expansions

The lines of disturbance diverge owing to

i. Increase of Mach number reducing Mach angle

ii. Flow turning towards source of disturbance.

c. Let the sources of disturbance 01 02 03.... approach each other at 0.

Figure 7.4 - ‘Expansion fan’ geometry

The divergence of the lines of disturbance ensures that the effect remains gradual. Expansion
takes place through a “fan” or Prandtl-Meyer expansion until the flow is parallel to surface S2.
Since each disturbance 0L1 0L2 ....0Ln is of infinitely small intensity the whole process is
isentropic.

d. The Prandtl-Meyer function relates the flow conditions before and after an expansion of the
type envisaged in (c) above. The standard analytical derivation is somewhat tedious and may be
obtained from any standard textbook if required. This will, in any case, be covered in detail in
the “Aerodynamics” notes and lectures on this Course. The theory allows for the calculation of

50
the fan angle and downstream conditions based upon knowledge of the incoming Mach number
and turn angle.
7.2 Flow round a concave corner

Consider in a similar manner the flow round a concave corner.

Figure 7.5 – Concave corner flow geometry

As in section 7.1, at the reference plane (1) the flow is parallel to surface S and at (2) parallel to
S2. Because area A2 is effectively less than A1 the flow decelerates.

u2 < u1 
Hence  i.e. a compression
p2 > p1 

a. Consider first an infinitely small compression originating at 0.

Figure 7.6 – Compression wave geometry

Behind the disturbance 0L1 the Mach number will be less, and hence the Mach angle greater than
before the disturbance. The flow will be turned away from the disturbance.

b. Now consider a series of small compressions originating at 01, 02 and 03 etc.

51
Figure 7.7 – Compression wave series geometry

The second disturbance 02L2 will converge on the first, joining forces with it at L12. The third
disturbance meets the resulting front at L23 and further disturbances augment the front further
out. A series of small compressions will always coalesce to form a shock front. (compare with
the expansion case). It should be noted that the second compression can only be formed in
supersonic flow and hence the first compression must take the form of a weak shock (the flow is
always subsonic after a strong shock - see section 5.1).

c. Let the sources of disturbance 01 02 03 etc approach each other at 0.

Figure 7.8 – Concave corner shock wave geometry

A single oblique shock is now formed at 0 which turns the flow so that it is parallel to S2. It

52
follows from (b) above that the oblique shock 0Lin. will be a weak shock. The process is non-
isentropic.
7.3 Detached Shocks

Consider the flow past a wedge, obtained by replacing the solid surface S1 of section 7.2 by a
streamline. Let the initial Mach number be fixed at some value M’. The flow pattern and its
location on a skeleton M1 - M2 chart for oblique shocks is given in Figure 7.9:

Figure 7.9 –Shock geometry from flow past a wedge

Suppose now the wedge angle δ is varied. The deflection may be increased from zero through
values δ1, δ2 etc until a maximum value of δ (δMAX) is reached, beyond which no solution is
possible, as shown on the skeleton diagram. The physical flow then changes its character. A
normal or near normal shock is formed ahead of the wedge. Behind this shock the flow is
subsonic and is therefore able to negotiate the wedge or body contour. The region of subsonic
flow is usually confined to the leading edge region, and away from this an oblique shock is
formed with supersonic flow both ahead of and behind the shock. The shock is curved as shown
in Figure 7.10. In order to obtain transition from the normal shock ahead of the body to the
oblique shock far distant from the body, the intermediate oblique shock must be a strong shock.

53
Figure 7.10 – Detached Shock geometry from flow past a wedge

The same phenomenon arises if the incident Mach number of an air-stream flowing past a fixed
wedge is varied. At high Mach numbers, two oblique shocks will be formed at the leading edge
of the wedge i.e. attached shocks. The flow configuration will correspond to that shown
overleaf. As the incident Mach number is reduced, the inclination of the shock will increase.
This can be seen from Figure 7.11, remembering that the deviation must remain constant at the
value corresponding to the wedge semi-angle as the Mach number is reduced. Numerical values
can be obtained from Figure 6.2 and Figure 6.3. At first the shock will remain attached, but the
Mach number can be reduced beyond a value for which there is a solution for the given value of
the deviation. The shock then becomes detached, and moves upstream as the Mach number is
further reduced.

Figure 7.11 – Shock geometry from a wedge over a range of Mach No

A bluff body is an extreme example of a wedge for which the local wedge angle is 90o. In that
case, a detached shock will be formed ahead of the body at all Mach numbers. The exact shape
of the shock and the distance it stands off from the leading edge depend on the geometry of the
body and on the approach Mach number. Due to the subsonic flow region, changes in body
geometry will be felt by the shock which must adjust itself to accommodate the flow

54
downstream. This contrasts with the usual supersonic flow situation. No analytical solution for
this type of flow is available, and empirical experimental data have to be used to determine the
flow pattern. Figure 7.12 illustrates these effects.

The detached shock formed ahead of a three-dimensional body follows essentially the same
pattern apart from the different numerical values which apply to the conical shocks set up in
these circumstances. It should be noted that since the maximum deviation for a cone is less than
that for a wedge, the detachment Mach number for a cone is less than that of a wedge of the
cone semi-angle.

Figure 7.12 – Detached shock geometries from a wedge and a bluff body

7.4 Multiple Shocks and Expansions

In this subsection the flow set up when there is more than one shock, or there is a shock and an
expansion, is examined qualitatively. The two particular cases quoted below are treated in a
general sense, but they occur frequently in practice either on their own or as constituent parts of
more complex flow patterns.

55
Figure 7.13 – Geometry of a shock followed by an expansion

a. Shock followed by expansion - For any shock the normal component of Mach number after
the shock is less than unity. Thus the angle β, which the streamline makes with the shock as
in the insert diagram below, is given by sin-1(Ma / M2) and is less than the Mach angle µ2
which is given by sin-1 (1 / M2). It follows that any disturbance, however slight or whatever
its sign, converges with the shock. This is illustrated in Figure 71.3 which shows a shock
followed by an expansion. The expansion generated at D interacts with the shock along BC,
reducing its strength and causing it to curve towards the local Mach line.

b. Shock followed by shock - If a shock is now generated at D, as in Figure 7.14 below, its
inclination ξ2 will be greater than µ2 and hence greater than β1. The shock DB then
converges on the shock 0B and they join to form the stronger shock BC. There must be no
pressure difference across the boundary, BF, between the flow affected by one shock and
that affected by two. Furthermore, the flow on each side of the boundary must be parallel to
the boundary. The one independent variable ξF, is not in general sufficient to satisfy both
conditions, so a disturbance BE is generated. The disturbance BE is usually of negligible
strength but is sufficient to satisfy the flow conditions by causing a small pressure change
and flow deviation, as at G. There is inevitably a discontinuity of velocity across the
boundary BF, which takes the form of a vortex sheet originating at the shock intersection B.

56
Figure 7.14 – Geometry of a shock followed by a shock

7.5 Shock Reflection

a. Reflection at a solid wall - A supersonic stream originally parallel to the solid wall BD is
deflected by a disturbance originating at A, as in Figure 7.15. In the case where deflection
occurs through the shock AB (left hand diagram), the flow is turned towards the wall. The
wall resists the turning and a reflected shock BC is generated to turn the stream parallel to
the wall. In the case where deflection occurs through the expansion fan A B1 B2 (right hand
diagram), the flow is turned away from the wall. In order to maintain contact the flow
undergoes the reflected expansion B1C1 B2C2.

Figure 7.15 – Geometry of a shock reflection off a solid wall

A special case arises when the flow is deflected by a shock if δ > δMAX for the particular

57
value of M2. The condition of zero δMAX deflection at the wall necessitates a normal shock as
shown in Figure 7.16. The initial shock is curved from B to B1, behind which the flow is
subsonic. B1C is the shock generated in the remaining supersonic flow. This condition is
called a Mach reflection.

Figure 7.16 – Conditions for generating a Mach reflection

b. Reflection at a free boundary - A parallel supersonic jet flowing in an atmosphere of


pressure p1 is deflected by a disturbance originating at A. In the case of the shock AB, which
meets the free boundary at B, the pressure p2 behind the shock is greater than p1. A free
boundary is unable to support this pressure and so an expansion fan B C1C2 is generated to
reduce the pressure to p1. In the case of the expansion A B1 B2, which meets the free
boundary over B1 B2, the pressure gradually diminishes below p1. A series of small shocks is
generated to restore the free boundary pressure to p1. They coalesce at C to form a strong
shock CD.

Figure 7.17 – Shock geometry from a reflection at a free boundary

c. Reflection at a plane of symmetry (free stream). - The conditions at a plane of symmetry are

58
somewhat different. Shown in Figure 7.18, two shocks AB and A1B meet at B. They deflect
the flow towards the plane of symmetry. At B, two mutually inclined streams impinge and
reflected shocks BC and BC1 are necessary to turn the two streams parallel to each other. A
plane of symmetry acts as a rigid boundary.

Figure 7.18 – Shock geometry from a reflection at a plane of symmetry

d. Shock Cancellation – Figure 7.19 shows the mechanism of deceleration of a supersonic flow
in a converging channel having straight walls. This only applies if the deflection is less than
δMAX for ME.

Figure 7.19 – Flow deceleration from shock cancellation

Figure 7.20 below shows the mechanism of expansion of a supersonic flow in a


divergent channel having straight walls.

Figure 7.20 – Flow acceleration from shock cancellation

59
Note that in both the above examples, the shocks and expansion lines cross the exit, giving a
flow varying in Mach number and direction. This may or may not be important. If
undesirable, the shocks and expansions must be cancelled. For example, the corner A
generates a shock reflected by the upper boundary at B. It is returned to the lower boundary at
C. If at exactly this point the lower boundary becomes parallel to the upper, the flow
deflection through BC is accommodated and no further shock is required (Figure 7.21).

Figure 7.21 – Shock cancellation from correct geometric matching

An example of cancellation of expansions is to be found in the nozzle. The converging portion


accelerates the flow to sonic velocity. The curvature of ABC generates expansions AD, BE and
CF. At the centre line these are reflected and return to the wall at G H and I. Along GHI the wall
must be so curved as to accommodate the flow deflection due to the expansion. Downstream of
FI the stream is then uniform in direction and Mach No.

Figure 7.22 – Shock cancellation inside an expansion nozzle

60
8. SHOCK—BOUNDARY LAYER INTERACTION

A detailed treatment of shock-boundary layer interaction is outside the scope of these notes,
though it is a subject of vital importance in the design of many propulsion systems. A very brief
physical description of the phenomenon is all that will be given. In section 6 it was assumed that
the flow was inviscid so that no boundary layers were present. In the actual case boundary layers
will be established on all duct walls. Since there is a static pressure rise across a shock the
boundary layer may be expected to behave as if a sharp pressure gradient were applied to it in
otherwise continuous flow. Two cases, of an initially laminar and an initially turbulent
boundary, are possible.
8.1 Laminar boundary layer
Although a pressure field is unable to propagate upstream of the shock in the supersonic
mainstream, it is possible in the case of a subsonic boundary layer. The first effect is therefore a
thickening of the laminar layer ahead of an impinging shock, brought about by the increasing
pressure in the direction of flow. This is illustrated in Figure 8.1.

Figure 8.1 – Laminar boundary layer thickening by shock interaction

The thickening of the laminar layer ahead of the main shock forms an effective, wedge just in
front of the foot of the main shock, and this wedge generates a second or oblique component to
the main shock which then joins the main shock some distance from the surface. This type of
shock formation is usually referred to as a bi-furcated or a lambda shock on account of its

61
appearance. Near the foot of the main shock, transition from laminar to turbulent flow will
occur. If the pressure rise is high, complete separation of the boundary layer from the surface
may take place, usually with undesirable consequences.
8.2 Turbulent boundary layer

If the boundary layer is already turbulent before the shock, only a small oblique component may
be formed. The boundary layer will thicken across the shock, as illustrated in Figure 8.2;

Figure 8.2 – Turbulent boundary layer thickening by shock interaction

If the shock is strong enough to cause local separation of the boundary layer, a bi-furcated shock
can be formed, as in the case of a laminar boundary layer. The patterns formed when an oblique
shock is reflected by a turbulent boundary layer are illustrated and compared with the no
boundary layer case in Figure 8.3:

Figure 8.3 – Shock geometry from turbulent boundary layer interaction

If a strong shock impinges on a turbulent boundary layer, complete separation can occur. It is
usually assumed that a static pressure ratio grater than 1.8 is likely to cause separation. Accepted

62
methods of boundary layer control, such as blowing or sucking can, of course, be used to
minimise separation. Practically, this involves some complication, and is difficult if the foot of
the shock moves along the wall with change of operating conditions.

The possibility of some shock boundary layer interactions of the nature just described should
always be considered when applying the shock formations outlined in section 7 to propulsion
components.

63
9. FLOW IN NOZZLES

A nozzle is used to produce the jet in all thermal propulsion systems. It is thus a component of
some practical importance as well as one illustrating a number of fundamental features of high
speed gas flow. The analytical treatment is usually based on the assumption that the flow is both
adiabatic and reversible. Although some heat transfer does take place from the gas to the nozzle
wall (and is of considerable importance in determining the nozzle wall temperature) the energy
so lost is negligible compared with the energy of the flowing gas and so the first assumption
closely represents the actual case so far as the gas flow is concerned. As regards the second
assumption it will be recalled that nozzles are usually plain ducts produced with a good surface
finish. Friction is therefore small and hence the second assumption closely represents the actual
case. Where friction is important it is usually taken into account by modifying the frictionless
results with empirical factors.
9.1 The convergent nozzle

The convergent nozzle is rarely used on missile propulsion systems. For the reasons given in
section 3.2 (iii), it cannot produce a supersonic jet, but on account of its simplicity could be used
in some subsonic applications. Let gas initially at pressure p1 and temperature T1 be expanded in
a convergent nozzle which discharges into an atmosphere of pressure pa as shown in Figure 9.1.

Figure 9.1 – Simple convergent nozzle geometry

When the. flow is subsonic the pressure at the nozzle exit (pe) will be equal to the atmospheric
pressure (pa) and, in the absence of losses, the total pressure will be equal to the initial pressure

64
(p1) assumed to be measured where the velocity is small. The ratio of total pressure to exit
pressure is given by:

γ (γ −1)
po  γ − 1 2  po po
= 1+ Me  = =
pe  2  pe pa

where Me is the Mach number in the nozzle exit plane.

Hence, as p1 / pa is increased, Me increases until it reaches unity, when

γ (γ −1)
po po  γ + 1 
= = (9.1)
pe pa  2 

The pressure ratio given by equation (9.1) at which sonic conditions are just reached at the throat
is often referred to as the critical pressure ratio. In the case of air, with γ = 1.4, the critical
pressure ratio (po / pe) = 1.893, i.e. if po / pe, the flow will be choked and Me = 1. The flow in that
case will be obtained by putting M = 1 in equation (2.10), i.e. by

−(γ +1) 2(γ −1) (γ +1) 2(γ −1)


m& To γ  γ + 1 γ  2 
= or m& = Ae po (9.2)
Ae po R  2  RTo  γ + 1 

If pa were decreased or p1 increased so that

γ (γ −1)
po  γ + 1 
> (9.1a)
pa  2 

supersonic flow cannot be established because there is no divergent section. In these


circumstances, pe can no longer be equal to the ambient pressure, but remains constant or
increases so that

γ (γ −1)
po  γ + 1 
= (9.1b)
pe  2 

A drop of static pressure from pe to pa takes place outside the nozzle, and is dissipated in noise
and turbulence, which also give rise to a loss in total pressure. It will be seen from equation (9.2)
that if the conditions represented by equation (9.1a) are achieved by reducing pa rather than by
increasing p1, the mass flow will remain constant, once the critical value of pe is reached at the
throat. Alternatively the mass flow may be expressed non-dimensionally as m& T1 At p1 and
plotted against the ratio p1 / pa as in Figure 9.2. The flat part of the curve, corresponding to a
constant mass flow when p1 & T1 are fixed, is often referred to as “choked” conditions.

65
The terms subcritical and supercritical are often used to describe non-choked and choked
conditions respectively.

Figure 9.2 – Mass flow – pressure ratio characteristic of a convergent nozzle

Example 6

An air-stream flows in a converging duct from a cross-sectional area of 0.001176 m2 to a


cross-sectional area of 0.001057 m2, where the pressure, temperature and velocity are
150 kPa, 125oC and 304 m/s respectively. (a) Determine the remaining fluid properties at
section 2, (b) determine the fluid properties at section 1, (c) show the process on a T-s
diagram, (d) find the mass flow rate.

Example 6 - Solution

(a) Using equation (2.5): a2 = √(γ R T) = √ (1.4 x 287 x 398.15) = 400 m/s
M2 = V2 / a2 = 304 / 400 = 0.76
Using equation of state: ρ2 = p2 / (R T2) = 150000 / (287 x 398.15) = 1.313 kg/m3

Using equation (2.7): To2 / T2 = 1.1155, ∴ To2 = 444.0 K

Using equation (2.8): po2 / p2 = 1.466, ∴ po2 = 219.9 kPa

Using equation (2.9): ρo2 / ρ2 = 1.3142, ∴ ρo2 = 1.726 kg/m3

(b) If the isentropic flow were maintained with the duct extended (as shown below) the
flow would be accelerated to M = 1, with the area being the critical area (A* or At).
Then, for M2 = 0.76:
Using equation (2.11): A2/A* = (1 / 0.76) [(2 + 0.4 x 0.762) / 2.4] (2.4/0.8) = 1.057

66
∴ A* = 0.001 m2 and A1 / A* = 1.176

From equation (2.11) or Fig 2.9, for A1/At = 1.057, M1 = 0.61 or 1.5 (impossible for
convergent nozzle).
For isentropic flow, the total properties do not change from one state to another, so:
To1 = 444.0 K, po1 = 219.9 kPa, ρo1 = 1.726 kg/m3

Using equation (2.7): To1 / T1 = 1.1155, ∴ T1 = 413.2 K

Using equation (2.8): po1 / p1 = 1.466, ∴ p1 = 171.0 kPa

Using equation (2.9): ρo1 / ρ1 = 1.3142, ∴ ρ1 = 1.442 kg/m3

a1 = √(γ R T) = √ (1.4 x 287 x 413.2) = 407.5 m/s, ∴ V1 = 0.61 x 407.5 = 248.6 m/s
(c) T – s diagram shown below.

(d) Using continuity equation, m& = ρ1 AV


1 1 = 1.442 x 0.001176 x 248.6 = 0.4216 kg/s

Example 7

A converging nozzle with an exit cross-sectional area of 0.001 m2 is operated with air at
a back pressure of 69.5 kPa. The nozzle is fed from a large reservoir where the stagnation
pressure and temperature are 100 kPa and 60oC respectively. Assuming steady, 1-D,
isentropic flow: (a) determine the Mach number and temperature at the nozzle exit, and
(b) the mass flow rate through the nozzle.

Example 7 - Solution

The first step is to check whether the converging nozzle is choking or not.

pb / po = 69.5 / 100 = 0.695 > p*/po (= 1 / 1.893 = 0.5283).

67
Therefore the flow is not choked and is in the subcritical flow regime.

Using equation (2.8): for pe / po = 0.695, Me = 0.74

Using equation (2.7): Te / To = 0.9013 or Te = 0.9013 x 333.15 = 300.3 K


Using equation of state: ρe = pe / (R Te) = 69500 / (287 x 300.3) = 0.8064 kg/m3

Using continuity equation,

m& = ρ1 AV
1 1 = 0.8064 x 0.001 x 0.74 x √(1.4 x 287 x 300.3) = 0.2073 kg/s

9.2 The convergent-divergent (con-di) or de Laval nozzle

Supersonic velocities can be achieved, when the overall pressure ratio is greater than critical, by
adding a divergent portion to the nozzle, as shown in Figure 9.3. The resulting nozzle shape is
the familiar convergent-divergent (sometimes called de Laval) nozzle, used on nearly all missile
propulsion systems.

Figure 9.3 – The convergent-divergent (de-Laval) nozzle geometry

If supersonic velocities are to be achieved in the divergent section, sonic conditions must be
reached at the throat (see section 4.2). It then follows that:

γ (γ −1)
 po  po  γ + 1 
 p  = p = 2 
 t t  

The mass flow will be obtained by putting M = 1 into equation (13), so that

(γ +1) 2(γ −1)


γ  2 
m& = At po (9.3)
RTo  γ + 1 

68
This is identical to the mass flow of a choked convergent nozzle. It depends solely on the
chamber pressure and temperature p1 and T1 and is independent of the ambient pressure. This,
however, is only true when sonic conditions exist in the throat. Although of no practical
importance for missile propulsion, it would be possible to establish a wholly subsonic flow, in
which case lower flows would be achieved.

The flow areas, the Mach numbers, and the velocities at each cross section can be calculated by
means of the equations derived in sections 2.5, 2.6 and 2.7, on the basis that the flow is
isentropic, i.e. po = constant. Analytical expressions are given in many text books, but are not
repeated in these notes since they have little practical value. Numerical values of the area ratio,
(A / At ) and the velocity ratio (u / ut ), have been plotted against pressure ratio (p1 / p) in Figure
9.4, where p is the pressure at any section. From a practical point of view, it should be noted that
the area ratio increases very rapidly with pressure ratio whereas the increase in velocity, which
determines the thrust is more moderate.

Example 8

A supersonic nozzle has throat and exit cross-sectional areas of 0.09 m2 and 0.15 m2
respectively. The mass flow rate of the air is 14 kg/s. The supply pressure and
temperature at the nozzle inlet, where the velocity is negligible, are 68 kPa and 40oC
respectively. Assuming steady, 1-D, isentropic flow: (a) determine the fluid properties at
the throat, and (b) determine the fluid properties at the exit of the nozzle.

Example 8 - Solution
The pressure and temperature at the nozzle inlet, where the velocity is negligible,
represent the total pressure and temperature, .i.e. po = 68 kPa and To = 313 K. Total
density (ρo) = po / (R To) = 68000 / (287 x 313.15) = 0.7567 kg/m3
(a) This is a supersonic nozzle so the flow at the throat must be sonic. Also the total
properties will be unchanged for in isentropic flow.

Using equation (2.7): for M = 1, γ = 1.4, To / T* = 1.20 ∴ T* = 260.8 K

Using equation (2.8): po / p* = 1.466, ∴ p* = 35.92 kPa

Using equation (2.9): ρo / ρ* = 1.3142, ∴ ρ* = 0.4797 kg/m3

V* = a* = √(γ R T) = √ (1.4 x 287 x 260.8) = 323.8 m/s

m& RTo
(b) The non-dimensional mass flow rate at the nozzle exit =
Ae po

69
= (14 x 287 x 313.15) / (0.15 x 68000) = 0.4115
Using equation (2.10) or Fig 2.8, Me = 1.98
Using equation (2.7): To / Te = 1.784 ∴ Te = 175.4 K

Using equation (2.8): po / pe = 7.587, ∴ pe = 8.962 kPa

Using equation (2.9): ρo / ρe = 4.2517, ∴ ρe = 0.178 kg/m3

Ve = 1.98 x √(1.4 x 287 x 175.4) = 525 m/s

Figure 9.4 – Convergent-divergent nozzle design parameters

Figure 9.4 also brings out an important feature. For a fixed value of γ there is a unique value of
(pa / p) for each value of the area ratio (A / At ). While these are the values that would be used
in design, it is clear that an existing nozzle may well have to operate at other values of (pa / p) in
some circumstances. No solution to these conditions is envisaged in any of the foregoing and so
“off design” operation of these nozzles must be examined in some detail. The flow regimes that
are set up in a convergent-divergent nozzle are best illustrated by considering the flow pattern in
a given nozzle as the ambient pressure (pa) is reduced from p1 to very low values. This is
relatively straightforward to set up experimentally, with a reservoir total pressure and the exit

70
(back) pressure (pe or pb) reduced via a valve.

Figure 9.5 – Convergent-divergent nozzle off-design flow regimes

i. If ( po pa ) < ( po pa )crit is less than the value of (po / pa) which first gives critical (sonic)

conditions at the throat, the velocity everywhere in the nozzle will be subsonic. Then,
γ (γ −1)
( po pt ) < (γ + 1) 2  . The pressure and Mach number distributions in the nozzle

are illustrated by the curves (1) in Figure 9.5. In this mode of operation, the gas
accelerates to the throat, but is still subsonic at that point. The throat is merely a turning
point and downstream of the throat the divergent portion of the nozzle acts as a diffuser,

71
so reducing the velocity and Mach number. In such cases the con-di nozzle is thus
behaving like a Venturi-meter. This condition of operation occurs only during start-up in
practical convergent-divergent nozzles.
γ (γ −1)
ii. If pa is reduced so that ( po pt ) = (γ + 1) 2  , i.e. ( po pa ) = ( po pa )crit , (= 1.893

for air) sonic conditions will be reached in the throat, the flow in the convergent section
being subsonic. The flow in the divergent section will also be subsonic. This is illustrated
by curve 2. If ( po pa ) > ( po pa )crit the flow will be supersonic in the divergent section.

It is important to note in order that the flow should be isentropic, two sets of equations
must be satisfied simultaneously. Curve 7 illustrates the flow conditions in this instance.
It is thus seen that two isentropic flow regimes are possible for the same flow up to and
in the throat. The one which is set-up will depend on the back pressure pa. That given by
the curve 2 is of little practical importance, but that given by the curve 7, in which
supersonic velocities are developed isentropically, represents the nozzle design operating
conditions.

iii. If pa lies between the values giving the flow regimes described at (ii) above, no
isentropic flow is possible.

iv. Remembering that if the flow is non-isentropic, there must be a loss in total pressure it
will be seen that if the ambient pressure is just below that defined by curve 2, a subsonic
exit velocity is possible. In these circumstances pe must be less than p1. Such a situation
could be created if a normal shock were situated at some point in the divergent portion of
the nozzle, as illustrated by the curves 3 in Figure 9.5. Indeed, this is the only mechanism
by which a flow meeting all the imposed conditions can be set-up, and is the one which
occurs in practice. In this mode of operation the fluid accelerates up to sonic velocity at
the throat and to supersonic velocities in the divergent portion after the throat. The
normal shock then restores subsonic velocities as illustrated. Since the working fluid has
expanded further than is necessary to give the outlet conditions the nozzle is said to be
over-expanded. It will be appreciated that the normal shock must occur at that point in
the divergent portion which enables all the flow conditions at the throat and exit to be
satisfied, the loss through the shock providing an additional variable for that purpose.

v. As the ambient pressure is reduced the normal shock moves downstream, as illustrated,
until it reaches the nozzle exit.

72
vi. Further reduction of the ambient pressure causes the shock system to move outside the
nozzle. Since a certain contraction of the stream is necessary (to correct the over-
expansion already noted in (iv) above) an oblique/normal shock system is set up. A fully-
developed oblique shock system is illustrated in Figure 9.5. The shocks are reflected as
shocks on the line of symmetry but as expansions where they strike the free jet surface,
so making the characteristic shock pattern.

vii. As the ambient pressure is further reduced the shock becomes more oblique, until the
shockless, isentropic regime 7 is established. This is, of course, the design condition
already referred to in (ii) above.

viii. If the ambient pressure is reduced beyond this value, insufficient expansion can take
place within the nozzle and further expansion now occurs outside the nozzle through an
expansion fan as illustrated. The nozzle is said to be under-expanded in this regime.

All the operating modes, from wholly subsonic to under-expanded supersonic, are shown in
sequences 1 to 8 in Figure 9.5. The graphs on the left hand side of the page gives the distribution
of pressure and Mach number within the nozzle for each case, while the small diagrams on the
right—hand side show the formation of shocks and expansions that will be set up. For a
completely expanded nozzle, each of the regimes 1 to 7 inclusive will be passed through during
start up, 7 being the final operating condition. Imperfectly expanded nozzles can correspond to
5, 6 or 8 in the operating condition.
9.3 Flow in actual nozzles

In practice, the gas flow in the nozzle will not correspond to the ideal for the following reasons:

i. Heat transfer to the nozzle walls, representing a direct loss of energy, though small. In a
regeneratively cooled nozzle this “loss” is made available to the system as a whole by
preheating of the fuel, and so does not constitute an over-all loss.

ii. A boundary layer will build up on the nozzle walls. Due to the accelerating nature of the
flow the boundary layer will be thin. When very high accuracy is required (e.g. when
nozzles are used for flow measurement) the boundary layer can be calculated by more or
less standard techniques and very good agreement is obtained. However, for all
propulsion applications, it is sufficient to allow for the boundary layer in the nozzle
throat by means of a discharge coefficient when high accuracy is required. In general it is
neglected.

73
iii. Imperfect flow distributions, i.e. non uniformity over any cross section. The gas velocity
is usually less near the walls. Velocity and temperature distributions are usually a result
of cooling techniques in the combustion chamber. Imperfect nozzle shaping may also
cause non-uniformities. This can change the line of thrust, which may be of more
practical importance than any losses involved.

iv. A uniform velocity is theoretically only possible with a contoured nozzle (see section
6.6(d)). The cost and difficulty of making such nozzles may lead to the adoption of a
conical supersonic divergence. It follows that the exit jet will have a radial component in
addition to the axial. The thrust from such a nozzle may be deduced by treating it as a
number of small concentric nozzles (See Figure 9.6).

Figure 9.6 – Integration variables for a conical nozzle

The axial component of thrust from each element will be given by:

dF = ρVe ( 2π r sin θ rdθ ) Ae cos θ + ( pe − pa )( 2π r sin θ rdθ ) cos θ

Integrating from 0 to α gives

1 + cos α
Fα = m& e + ( pe − pa ) Ae or Fα = λ Fo
2

1 + cos α
where λ = , Fo = thrust from a nozzle giving a parallel exit flow.
2

Values of λ are tabulated below:

α0 0 5 10 15 20 25 30

λ 1.0 0.998 0.992 0.983 0.970 0.953 0.933

74
v. Continued burning of the fuel and oxidiser may occur in the nozzle. Also recombination
of dissociated gases can take place in the nozzle. Since these both lead to an increase in
the kinetic energy of the gases they increase the nozzle efficiency. Practical calculations
are usually carried out on the basis of either continuous or frozen equilibrium as
discussed in section 1.l. Generally speaking the residence time is so short that frozen
conditions apply, but this could not be so in some advanced propulsion proposals.

vi. Shock waves which occur in an over-expanded nozzle can cause separation of the flow
from the walls by one of the mechanisms described in section 8. This may solely reduce
the efficiency. If the degree of over-expansion is great so that a strong normal shock is
formed, the location of the separation zone can be unstable and give rise to unacceptably
large fluctuating forces. To avoid flow separation, the pressure ratio (po / pa) should not
be less than about a third of its design value.

vii. It should be noted that normal shocks in the divergent section are found only in air
nozzles or channels with small angles of divergence. Normal shocks are not found at all
in rocket nozzles.

75
10. GAS FLOW WITH HEAT TRANSFER

Only the simplest flow with heat addition (or subtraction) will be considered. The treatment will,
however, cover practical applications. It is limited by the following conditions:-

a. The area of the duct is constant.

b. The flow is steady and one-dimensional.

c. The rate of heat addition (or removal) at any section is constant.

d. Effects of friction are negligible.

Assumption (d) means that either heat transfer effects are so great as to swamp friction, or that
the duct length is so short that friction is negligible, e.g. in extreme case, detonation.
10.1 Relationship between flow before and after heat transfer

With these assumptions the analytical treatment is straightforward. Applying the energy,
momentum and continuity relations we get:

a. Energy:

q = c pTo 2 − c pTo1

b. Momentum:

For a parallel frictionless duct, F%2 = F%1 or po 2G2 = po1G1

c. continuity:

Apo1
m& = X 1 (γ , M )
RTo1 To1  po1  X 1 (γ , M )
i.e. = 
Apo 2 To 2  po 2  X 2 (γ , M )
= X 2 (γ , M )
RTo 2

Substituting for (po1 / po2) gives:

To1 G2 X 1 X 1 G1 X 2 G2
= or =
To 2 X 2G1 To1 To 2

Substituting for G1 and G2, this may be re-written as:

N1 N
= 2 (10.1)
To1 To 2

76
 γ −1 2 
M 1 + M 
 2 
where N= (10.1a)
γ M +1
2

Both G and N are presented graphically on page plotted against Mach number. These enable the
flow conditions before and after the heat transfer to be calculated.
10.2 Physical Limitations on flow

Writing equation (10.1) as N 2 = N1 To 2 To1

It is seen that heat addition will give an increased value of N, and hence an increase in the Mach
number of the gas stream, if the flow is subsonic, or a decrease in the Mach number if the flow
is supersonic. Heat removal would have the opposite effect. From the graph (or by
differentiation of (10.1a) it can be shown that N has a maximum at M = l.

Hence heat addition will cause the Mach number of any gas stream to approach unity. When
sufficient heat has been added to give unity Mach number, a condition known as thermal
choking is obtained. Any further addition of heat must cause a change in the upstream conditions
of the flow.

It should also be noted that the function G also has a maximum at M = 1.0. Heat addition will
therefore cause an increase in G and hence a reduction in the total pressure. This is sometimes
referred to as the “fundamental pressure loss (fpl)” since it occurs even in the absence of
friction from walls, flame stabilisers etc. The fpl is physically due to a static pressure difference
being necessary to provide the force to accelerate the flow, i.e. to satisfy the momentum
equation.

The consequences of heat addition to a gas stream can therefore be summarised as:

i. Addition of heat to a subsonic stream:

a. Static pressure, total pressure, and density decrease.

b. Mach number increases until M = 1 .0 when thermal choking occurs.

ii. Addition of heat to a supersonic stream

a. Static pressure and density increase

b. Total pressure and Mach number decrease until M = 1 when thermal choking occurs.

It should be noted that a more rigorous treatment of the above phenomenon can be obtained by

77
 T  p (γ −1) γ 
expressing the entropy change as: ∆s = c p log e  2  1  
 T1  p2  

Substituting for the static temperature ratio and for the static pressure ratio, the right hand side of
equation can be expressed solely in terms of γ and M as in equation (10.2) below:

(γ +1) γ

∆s = c p log e 
2 (
 M 2 1+ γ M 2
1 ) 
 (10.2)
 2 ( γ +1) γ 
 1 (
M 1+ γ M 2
2
) 

This has a maximum, and only one maximum in terms of M2 when M2 = 1.0.

Likewise the enthalpy ratio can be expressed solely in terms of γ and M as in equation (10.3)
below

( )
2
M 2 1 + γ M1
2 2
H2 c pT2
= = (10.3)
H1 c pT1 M 2
1 (1 + γ M )
2
2 2

This has a only one maximum, when M 2 = 1 γ .

Possible states of the gas can be represented on the entropy-enthalpy chart by the curve
(Rayleigh line) shown in Figure 10.1:

Figure 10.1 – Heat transfer enthalpy – entropy characteristics for a supersonic flow

78
The limitations of a heating or cooling process as discussed above and some of the consequences
are clearly brought out in Figure 10.1. In particular it should be noted that the attainment of
sonic conditions, whether by heating on the subsonic branch or heating on the supersonic
branch, coincides with maximum entropy.

Figure 10.2 – Mach No – N function characteristic for a range of G

79

You might also like