Download as pdf or txt
Download as pdf or txt
You are on page 1of 130

AEROSPACE PROPULSION SYSTEMS

Courtesy NASA
Courtesy USAF

Chapter 1 Fundamentals
Chapter 2 Rockets
Chapter 3 Piston Aerodynamic Engines
Chapter 4 Gas Turbine Engines
Chapter 5 Ramjets and Scramjets
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd
Aerospace Propulsion Systems

CHAPTER 1 - FUNDAMENTALS

“Give me six hours to chop down a tree and I will spend the first four sharpening the axe.”
– US President Abraham Lincoln
WHY STUDY PROPULSION?

 From the earliest days


of recorded history,
many have dreamed of
soaring into the sky.
 Aerospace propulsion
systems are the means
for attaining powered
flight.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-3
WHY STUDY PROPULSION?

 The technology has allowed the peoples of the


world to be drawn closer together with
commercial air transportation …

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-4
WHY STUDY PROPULSION?

… and has expanded our frontiers into space.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-5
Propulsion

 Propulsion means to drive forward. Therefore a


propulsion system is a machine that produces
a thrusting force to drive an object forward.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-6
Thrust

 Most aerospace propulsion systems produce


thrust (FN) by applying Newton’s Third Law of
action/reaction.
 Thrust is produced by accelerating a working
gas (normally by adding heat due to chemical
combustion).

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-7
Types of Aerospace Propulsion Systems

 Rockets
• Oldest type, dating back to the
Han Dynasty in China (circa
1,000 BC).
• Used today in space launchers
and missiles.
• Can operate outside Earth’s
atmosphere.
• Capable of very high thrusts. Courtesy National Museum of the
USAF

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-8
Types of Aerospace Propulsion Systems

 Piston Aerodynamic Engines


• Used in general aviation aircraft.
• Relatively low cost
• Capable of low thrusts
• Limited to low subsonic speeds
Courtesy National Museum of the
• Limited to low altitudes USAF

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-9
Types of Aerospace Propulsion Systems

 Turbojet Engines
• The core of all gas turbine
engines.
• No longer used in aircraft, but
still used in missiles.
• Capable of high thrusts. Courtesy NASA

• Capable of supersonic speeds


with the use of an afterburner.
• Poor fuel efficiency
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-10
Types of Aerospace Propulsion Systems

 Turbofan Engines
• Widely used today in commercial
and military aircraft.
• Capable of high to medium
thrusts.
• Capable of supersonic speeds
(typically requires an afterburner).
• Better fuel efficiency than
turbojets.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-11
Types of Aerospace Propulsion Systems
 Turboprop engines
• Used in short-range commercial
aircraft and military
transports/cargo aircraft.
• More fuel efficient than
turbofans.
• Limited to medium altitudes Courtesy National Museum of the USAF

and subsonic speeds.


• Short take-off and landing
• Noisy, vibration
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-12
Types of Aerospace Propulsion Systems

 Turboshaft Engines
• Used in helicopters and
in auxiliary power units
(APUs).
• Optimized to produce Courtesy USAF

shaft power.
• Generally short in length.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-13
Types of Aerospace Propulsion Systems
 Ramjet Engines
• Used in long-range supersonic
missiles and some specialty
aircraft.
Courtesy NASA

• Mechanically simple (few moving


parts).
• Can operate efficiently at Mach 2.5
to 5.0.
• Most cannot operate at subsonic
speeds, so require a booster rocket.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-14
Types of Aerospace Propulsion Systems

Courtesy NASA

 Scramjet Engines
• Used in experimental hypersonic vehicles.
• Many difficult technical challenges. No operational
models yet.
• Can operate at hypersonic Mach numbers 5.0 to
15.0.
• Cannot operate at low supersonic or subsonic
speeds, so it requires a booster rocket.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-15
1.1a Review of Terms

 System
• An identifiable collection of matter that is under
investigation.
 Types of Systems
• Isolated – Uninfluenced by its surroundings.
• Closed – Contains a fixed mass.
• Open (or Flow System) – Mass can transfer across
the boundary of a control volume.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-16
1.1a Review of Terms

 Working fluid
• Most aerospace propulsion systems operate in a
thermodynamic cycle involving transferring heat to
and from a working fluid.
• Generally this is atmospheric air (or air mixed with
combustion gases).
• The properties of air (e.g. density, temperature,
pressure, etc.) change with altitude. This limits the
altitude and speed (flight envelope) of air-breathing
engines.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-17
Flight envelopes of different aircraft

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-18
1.1a Review of Terms

 Work (W) is a form of


energy generated
when a force moves
something in the
direction it is being
applied.

The definition of work requires movement to occur. The girl in this figure makes a
great point, but she obviously did not see her dad typing on the keyboard, moving
his mouse or putting paper in the printer.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-19
1.1a Review of Terms

 Power ( W ) is the rate of doing work.


 Energy (E) is simply the capacity to do work and
exists in many forms (e.g. kinetic, potential,
thermal, mechanical, electrical, chemical,
magnetic, or nuclear).

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-20
1.1a Review of Terms
 Heat – The energy transferred between
molecules of one system to another due to a
temperature difference.
 Heat transfer mechanisms
• Conduction – Takes place between two motionless
adjacent substances at different temperatures
• Convection – Takes place between a solid surface
and a fluid in motion adjacent to it.
• Radiation – Energy emitted by electromagnetic
waves. This can occur across a vacuum (space).
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-21
1.1a Review of Terms
 Thermal efficiency (h th)
• A performance measure of merit of a heat engine
• The ratio of its work output to the total heat added
into the system:
• The 2nd Law of Thermodynamics states that ηth
can never reach 100%.

W Qin  Qout Qout


 th    1
Qin Qin Qin
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-22
1.1a Review of Terms

 Process
• A system undergoes a process when its state
changes from one equilibrium condition to another.
 Cycle
• If a system undergoes a number of processes so
that its final state equals its initial state, then the
system has undergone a cycle.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-23
1.1a Review of Terms
 This figure shows an ideal gas
turbine engine cycle.
• Process 4→1 is listed as “Not
Possible” because it occurs
outside the engine when the
exhaust gases diffuse in the
ambient air surroundings.
• However since this is an open
system, with new intake air
continuously flowing into the
engine (restarting the cycle at
point 1), it effectively is a cycle.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-24
1.1a Review of Terms

 Reversible processes
• The original state of the system can be restored,
leaving no change in the surroundings. This is an
ideal process which does not occur in nature.
 Adiabatic
• A process in which no heat transfer occurs.
 Isentropic
• A reversible, adiabatic process.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-25
1.1a Review of Terms
 Entropy (S)
• A measure of disorder, chaos, or randomness. For
an isolated system:
ΔS > 0 Irreversible processes
ΔS = 0 Reversible (isentropic) processes
ΔS < 0 Impossible

• Isentropic efficiencies will later be used to compare


the actual performance of engine components to
their idealized performance.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-26
1.1a Review of Terms

The 2nd Law of Thermodynamics shows that processes will naturally become more
disordered. By intentionally adding heat or work to a system we can bring more
order to a particular attribute of the system. However, even though one attribute of
the system may become more ordered, there will always be a net entropy gain (or
increase of disorder) due to other irreversibilities (such as friction or heat loss).

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-27
1.1b Perfect Gas Law (Equation of State)
 The properties of a working fluid (like air) are very
important in analyzing propulsion systems.
 Air is closely approximated as an ideal gas. Therefore
the equation of state for a perfect gas is a useful
analysis tool.

Pv  RT

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-28
1.1b Perfect Gas Law (Equation of State)

 In the previous equation, R is the gas constant.


It is related to the universal gas constant (Ro)
by:

o Ro
R ≡ Universal gas constant [=8.3145 kJ/(kmoles∙K)]
R
Mw Mw ≡ Molecular weight of gas [kg/kmoles]

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-29
1.1c Conservation of Mass

 

t control
 dV    V  dA  0
control
volume surface

For steady flow:    V A  c o n s t a nt


m

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-30
Example 1.1
Liquid oxygen (LOX) and liquid
hydrogen are steadily injected into
a rocket thrust chamber at 8 kg/s
and 1 kg/s respectively and
ignited. The combustion products
are expelled from the rocket
through a nozzle with a diameter
of 25 cm. If the density of the
combustion gases is 0.175 kg/m3.
Determine the exit velocity of the
combustion gases.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-31
Example 1.1

Solution
The control volume of the rocket thrust chamber and nozzle is shown
by the dotted line in the figure. Since the propellants are flowing at a
steady rate, the conservation of mass equations are reduced to:

0, steady flow
 

t control
 dV    V  dA  0
control
volume surface

 comb Ae Ve  m
O  m
H 2 2
prod
 
 exit
m  in
m
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-32
Example 1.1

m O2  m H 2 m O2  m H 2
Ve  
Ae  comb  d2  
prod   comb 
4  prod 
8 kgs  1 kgs
  1,047.7
 
m

4  0.25 m 
 2 kg s
0.175 m3

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-33
1.1d Conservation of Momentum

 
   
F  
t control
 V dV   V  V  dA
control
volume surface

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-34
Example 1.2

A rocket motor burning on a test stand steadily exhausts 20 kg/s of


combustion gases at an exit velocity of 750 m/s. The static pressure
of the exhaust gases exiting the nozzle is 120 kPa. Assume an
ambient air pressure of 101.3 kPa. Determine the force (or thrust)
the rocket produces.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-35
Example 1.2
Solution
 The external reaction force (Rx) which holds the rocket in place on the
test stand is equal in magnitude but opposite in direction to the
thrust force produced by the rocket

Rx   FN
 The control volume encompassing the rocket and test stand is shown
by the dotted line in the figure. Since the exhaust gas is flowing at a
steady rate, the conservation of momentum equations reduce to:
0, steady flow
 
   
FN  Pe  P0  Ae    V dV  V  V  dA
t control control
Aerospace Propulsion Systems Thomas A. Ward volume surface
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-36
Example 1.2
FN  m V  Ae Pe  P0 


 20  750 
kg
s
  0.02 m  120 kPa  101.3 kPa 
m
s 2
N
1,000 kN

 15.4 kN

This example shows how the thrust equation for a rocket is


derived from the Law of the Conservation of Momentum.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-37
1.1e Conservation of Energy

2
  
d 
 Q  W    e dV     u   g z   V  dA
t control
V
 
control  
dt 2
volume surface

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-38
1.1e Conservation of Energy
 The specific enthalpy (h) is defined as:
P
hu

 Substituting this into the energy equation gives an
alternative form:
2
  
d 
Q  W '   e dV   
t control
 h 
V

 g z   V  dA
dt control  2 
volume surface

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-39
1.1e Conservation of Energy

 An important parameter in this equation is the


specific heat, defined as the amount of heat
required to raise a unit mass through a 1 ºC
temperature rise.
 u 
• When volume is held constant: Cv 
 T 
V

• When pressure is held constant: C   h 


P  T 
P
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-40
1.1e Conservation of Energy

 The gas constant (R) is related to these two


specific heats by:
dh du
R   C P  Cv
dT dT

 Also a ratio of specific heats ( ) can be defined


as:
CP

Cv
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-41
1.1e Conservation of Energy

 A calorically perfect gas is one that has


constant specific heats. If such a gas is in a
thermodynamic process between two states,
then the definitions for Cv and Cp can be
integrated to obtain:
2
 u  u 2  u1   Cv dT  Cv T2  T1 
1

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-42
1.1e Conservation of Energy

 And …
2
 h  h2  h1   C P dT  C P T2  T1 
1

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-43
Example 1.3
Air enters an adiabatic compressor
of a turbojet engine at 70 kg/s and
at a temperature (T1) of 30°C. It
flows steadily through the
compressor with no change in
velocity and exits at a temperature
(T2) of 350°C. Assume that the
constant pressure specific heat (Cp)
of air is 1.005 kJ/(kg·K). Determine
the minimum power that must be
generated by a turbine in order to
drive the compressor at these
conditions.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-44
Example 1.3

Solution
The control volume around the compressor section is shown by
the dotted line in the figure. The energy equation for steady
flow is:
No velocity change
0, steady flow No height change 0, adiabatic
2
  


t control
e dV     h 

V

d

 g z   V  dA  Q  W ' 
control  
2 dt
volume surface

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-45
Example 1.3
Therefore the power required by the compressor to pressurize the
air (and thereby also increase its temperature) is equal to the
minimum power required by the turbine to drive it.

T1  30C  273  303 K


T2  350C  273  623 K

Wturbine  W compressor  m air  h2  h1   m air C p  T2  T1 


 
 70 kgs 1.005 kgkJK  623 K  303 K 
 22,512 kJs or kW
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-46
1.2 Isentropic Equations
 Isentropic equations are used often in the analysis of gas
turbine engines. (Standardized tables show the isentropic
parameters.)

 These are derived by first applying the 1st Law of


Thermodynamics to an isentropic process, which gives:

dQ  dW  dU
T dS  P dV  Cv dT
 P dV
dT 
Aerospace Propulsion Systems Thomas A. Ward
Cv 47
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-47
1.2 Isentropic Equations
Recall for a perfect gas:

Pv
T
R
Substituting this into the equation yields:

dT  P dv  v dP 
1
R

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-48
1.2 Isentropic Equations
Differentiating gives:

 P dv  v dP 
P dv 1

Cv R

Cv
P dv  v dP   P dv  0
R

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-49
1.2 Isentropic Equations
Also recall:
Cv 1

R  1

Substituting this in the equation gives:

1
P dv  v dP   P dv  0
 1

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-50
1.2 Isentropic Equations

P dv  v dP    1 P dv  0

P dv  v dP   Pdv  P dv  0

dP dv
 0
P v
ln P   ln(v)  0
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-51
1.2 Isentropic Equations
Applying logarithmic identities transforms these
equations into:

 
ln P   ln v  ln P v  0

 

Therefore:

P v  c o n s t a nt  C

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-52
1.2 Isentropic Equations
Again recall that for a perfect gas:

RT
v
P
Substituting this in gives:

  
 RT 
 T R
Pv P    1  c o n s t a n t  C
 P  P

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-53
1.2 Isentropic Equations
   
T1 R T2 R
 1   1
P1 P2
This results in the following isentropic equation that
relates temperature and pressure:
 1
T1  P1  
 
T2  P2 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-54
1.2 Isentropic Equations
 The gas turbine engine cycle is
commonly illustrated by a T-S
diagram of the engine cycle.
• The ideal (or isentropic)
processes representing the
compressor and turbine are
shown by the vertical arrows and
labeled with a prime symbol (′).
• The actual (or non-isentropic)
processes are represented by
dashed arrows
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-55
1.2 Isentropic Equations
It is also useful to develop isentropic expressions that
relate the specific volume ( v) to pressure (P) and
temperature (T). Starting with the definition of
enthalpy (h):

P
dh  du  d    du  d P v 

dh  du  P dv  v dP
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-56
1.2 Isentropic Equations

Recall:
du  Cv dT   P dv

Substituting the definition of enthalpy (h) into


this equation gives:

0  (dh  P dv  v dP)  P dv

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-57
1.2 Isentropic Equations

0  dh  v dP
dh  v dP
Since:
dh  C p dT
then:
v dP  C p dT

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-58
1.2 Isentropic Equations
v dP  P dv
 Since: dT  
Cp Cv

 P dv Cv

v dP Cp

dP  C p dv dv
  
P Cv v v
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-59
1.2 Isentropic Equations
Integrating this equation between points 1 and 2 gives:

P2 v2
dP dv

P1
P
  
v1
v

 P2   v2 
ln     ln  
 P1   v1 

 P2   v2 
ln    ln  
 P1   v1 
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-60
1.2 Isentropic Equations
 Resulting in an isentropic equation relating pressure
and specific volume:

 
P2  v 2   v1 
     
P1  v1   v2 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-61
1.2 Isentropic Equations
 Again recall for a perfect gas:
RT
P
v
 This can be substituted in to give a relationship with
temperature.
R T2

P2 v2  v1 
   
P1 R T1  v2 
v1
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-62
1.2 Isentropic Equations

T2  v1   v1 
    
T1  v2   v2 
1 
T2  v1   v1 
    
T1  v2   v2 
 Resulting in an isentropic equation relating
temperature and specific volume:
 1
T2  v1 
  
Aerospace Propulsion Systems Thomas A. Ward
T1  v2 
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-63
1.3 Polytropic Processes

 Polytropic processes follow laws that form the


relation:
P vn  c o n s t a nt  C

Where 0 < n < 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-64
1.3 Polytropic Processes

 Isentropic processes can be thought of simply


as a polytropic process with n =  .
 This offers an alternative method from
isentropic processes of defining efficiencies.
 However for simplicity, this presentation will
deal only with isentropic efficiencies

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-65
1.4 Total properties

 When a fluid in motion is isentropically brought


to rest a temperature and pressure rise occurs.
The fluid properties at this point are known as
stagnation properties.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-66
1.4 Total properties

 In the absence of hydrostatic pressures (e.g.


elevation effects of fluid weight on pressure),
stagnation properties are equivalent to total
properties.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-67
1.4 Total properties

 It can be shown by applying the energy


equation that the total temperature (Tt) is:

Total temperature = Static temperature + Dynamic temperature

2
V
Tt  T 
2C p

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-68
Mach Number (M)

 It is often convenient to describe these values


in terms of Mach number, rather than velocity.
 The Mach number (M) is non-dimensional
number defined as the ratio of the velocity (V)
over the speed of sound (a):

V
M 
a
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-69
Speed of Sound (a)

 The speed of sound for a perfect gas (such as


air) is:
P
a   RT

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-70
1.4 Total properties
 Substituting these definitions into the equations for
total pressure (Pt) and total temperature (Tt) gives:

Tt
 1
M 2
  1 
T 2

Pt  M 2   1  1
 1  
P  2 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-71
X-Function

 Like Mach number there are several other non-


dimensional parameters that are useful for
propulsion analysis. One commonly used
parameter is simply known as the X-function.
 It is derived by substituting the Equation of
State into the steady flow Conservation of Mass
equation:
P AM 
m 
RT
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-72
X-Function

 By substituting the definitions just derived for


Pt and Tt, this equation becomes:

   1 

m 
Pt A M  
1
   1 M 2  2   1 

 
R Tt  2 

 A portion of this equation is the X-function.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-73
X-Function

 The X-function (or the non-dimensional mass


flow rate) is defined as:
  1
  1 2  2  1
X  M  1  M 
 2 

 Substituting this back into the previous mass


flow equation, gives:
m R Tt
X
Aerospace Propulsion Systems Thomas A. Ward
A Pt
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-74
Y and Z-Functions

 The X-function can be further segmented into


the Y and Z-functions:
X Y Z
 The Y-function (or the non-dimensional specific
internal thrust reciprocal) is:

  1  2
M  1  M
 2 
Y
1  M 2
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-75
Y and Z-Functions

 The Z-function (or the non-dimensional internal


thrust) is:

Z
1   M 
2


   1 2   1
1  2 M 

 The utility of these functions will become


evident later

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-76
1.5a Isentropic Flow in Ducts

 Steady, isentropic flow through a frictionless


duct is one of the simplest fluid dynamics
systems to define and analyze.
• No work can be extracted from a duct.
• If the flow is isentropic (adiabatic) then there is no
heat transfer (in or out), therefore the energy
equation is:

Q W  0

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-77
1.5a Isentropic Flow in Ducts

 Since there is no heat transfer then:  Tt  0


 For steady flow (  m  0) in a frictionless duct
then also:  Pt  0
 Had friction losses been considered (non-
isentropic), there would be a loss in total
pressure ( Pt 2  Pt1 ). This is evident by inspection
of the following equation:
S  Pt 2 
s    R ln  
Aerospace Propulsion Systems Thomas A. Ward
m  Pt1 
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-78
1.5a Isentropic Flow in Ducts

 The continuity equation for one-dimensional


steady flow through a varying differential
control volume can be written as:

  d  A  dAV  dV   0
d dA dV
  0
 A V

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-79
1.5a Isentropic Flow in Ducts

 Applying the linear momentum equation for this


same control volume gives the equation:
 dP 
p AP   dA  P  dP  A  dA   AV dV
 2 
dP   V dV  0
 Combining these equations:
 d dA 
dP   V 
2
 0
  A
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-80
1.5a Isentropic Flow in Ducts
 P 
 But since: a   
2

   s

 Substituting this in gives:


 dP dA 
dP   V      0
2

 a
2
A

dP 1  M  2
  V
2 dA
A

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-81
1.5a Isentropic Flow in Ducts

 This equation is significant because it shows


the effect that Mach number has on flow inside
a varying area duct or channel.
Duct Geometry Entry Mach Static Velocity
No. Pressure
Converging M<1 Decreasing Increasing
(Decreasing area)
M>1 Increasing Decreasing
Diverging M<1 Increasing Decreasing
(Increasing area)
M>1 Decreasing Increasing
Converging-Diverging M<1 Decreasing Increasing
(Decreasing-then-Increasing)

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-82
1.5b Turbomachinery

 Turbomachinery (such as compressors or


turbines) are designed to transfer work (but not
heat).
• Compressors are used to increase the pressure of
a flow.
• Turbines are used to extract work (or energy) from
the flow.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-83
1.5b Turbomachinery
 The power required to drive a compressor is
determined by the energy equation as:
 C p Tt 2  Tt1 
Wc  m
 Ideally, the temperature ratio in compressors and
turbines would be the minimum associated with
pressure changes. So the process is isentropic, and
 1
Tt1  Pt 2  
 
Tt 2  Pt1 
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-84
1.5b Turbomachinery

 However, in reality there are irreversibilities due


to friction on all of the large wetted surfaces.
 Entropy does increase but not used as a
normal basis for analysis. Instead isentropic
efficiencies are used as the ideal against which
the actual is rated.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-85
1.5c Combustion Chambers

 Combustors are designed as steady flow


devices. They are essentially ducts with the
capacity for heat addition. No work can be
extracted.
 The heat produced by a combustor is:
 
Q  m C p Tt 2  Tt1 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-86
1.5d Nozzles

 Two types of nozzle


shapes are primarily
used in aerospace
propulsion systems:
• Convergent nozzles
• Convergent-divergent
(condi) nozzles

• A 3rd type, divergent nozzles, are used in scramjets.


Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-87
1.5d(i) Convergent Nozzles
 Isentropic flow through a convergent
nozzle can best be understood by
examining a nozzle with a constant
chamber pressure (Pc) and applying
decreasing back pressures (points A→D)
on it.
 If the nozzle is exhausting gases into the
atmosphere, this back pressure is equal
to the ambient pressure (P0). Therefore a
continuous decrease in the ambient back
pressure is equivalent to climbing in
altitude
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-88
1.5d (i) Convergent Nozzles
 At Point A: P0 = Pc, so there is no mass
flow through the nozzle. As P0 is lowered
to point B and beyond, the static pressure
through the nozzle decreases and the
mass flow through the nozzle increases.
The Mach number at the nozzle exit plane
also increases. Under these conditions,
the exit static pressure (Pe) is equal to P0.
 This is called fully expanded flow. This is
an optimal condition for propulsive
convergent nozzles, because it maximizes
the net thrust.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-89
1.5d (i) Convergent Nozzles
 This trend continues until point C is
reached, where the fluid exiting the
nozzle is equal to the velocity of sound
(Me=1.0 ).
 Since flow through a convergent nozzle
cannot be accelerated from subsonic
velocities to supersonic velocities;
therefore as the back pressure continues
to decrease past point C (to point D and
beyond) no additional mass can flow
through the nozzle. This is called choked
flow.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-90
1.5d(i) Convergent Nozzles
 For choked flow, the exit static
pressure is not equal to the back
pressure (Pe ≠ P0).
 Under these conditions the sonic
gases will dissipate through a
shock system. If the pressure
differences are large enough,
these shocks will form outside
the nozzle.
Choked flow means that the mass flow rate has
reached a maximum value.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-91
1.5d (i) Convergent Nozzles
 Choked (or underexpanded) flow occurs when the pressure
ratio ( Pte P0 ) is greater than or equal to a critical value. This
places a limit on the air mass flow that can flow through the
nozzle
 The optimum (or maximum) thrust occurs when the exhaust
gases fully expand to the ambient pressure (Pe = P0). This
maximizes the momentum thrust.
 Choked flow results in a loss of momentum thrust, but creates
a smaller pressure thrust component since (Pe > P0). This lost
momentum thrust may only be recovered by adding a divergent
surface (e.g. condi nozzle).
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-92
Critical Pressure Ratio

 For isentropic flow, an equation can be derived


that defines the critical pressure ratio (PRcrit)
necessary to just choke the nozzle.
• This is the maximum pressure ratio (Pt /P) that can
be achieved in the nozzle. This will occur when
Me = 1.0 (for a convergent nozzle).

Pt    1   1
PRcrit   1  
P  2 
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-93
Choke Test
 The critical pressure ratio is a function of γ only (e.g. for
γ = 1.333 then PRcrit = 1.852 and for γ = 1.4 then
PRcrit = 1.893).
 It provides a test to see if a convergent nozzle is choked or not.
• If the ideal pressure ratio (achieved by full expansion of the
flow through the nozzle) exceeds the critical pressure ratio
than this ideal ratio cannot be achieved because flow
through the nozzle is choked.

Pte
Choked if:  PRcrit
P0
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-94
Choked Flow
 If the nozzle is choked, then the exhaust pressure
ratio (Pte/Pe) is equal to the maximum or critical
pressure ratio (PRcrit). Therefore the static pressure of
the exhaust gases is:

 Pe  Pte
Pe  Pte   
 Pte  PRcrit

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-95
Unchoked Flow
 If the nozzle is not choked, flow is subsonic
throughout the nozzle (Me<1).
 Flow through the nozzle can adjust to changes in
ambient back pressure (altitude). Ambient pressure
changes will propagate upstream from the nozzle
exhaust plane at the speed of sound.
 So for all unchoked flows in a convergent nozzle, the
exit pressure will be equal to the ambient back
pressure (Pe=P0).

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-96
1.5d (i) Convergent Nozzles
 The modes of operation of a convergent nozzle are summarized
in the following table:
Modes of Operation – Convergent Nozzle

Under-expanded Fully expanded


Just choked
(choked) (not choked)

Pte P0  PRcrit PRcrit  PRcrit


Pte Pe PRcrit PRcrit  PRcrit
Me 1 1 <1
Pe  P0 P0 P0
Ae Pe  P0  >0 0 0
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-97
Critical Temperature Ratio
 An equation for the critical temperature ratio for
choked flow can also be determined:
 1
TRcrit 
2
 The exit static temperature for a choked nozzle can be
determined from the critical temperature ratio:

Tt
Te 
TRcrit
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-98
1.5d(ii) Convergent-Divergent Nozzles
 As was done for convergent nozzles,
isentropic flow through a condi nozzle can
be understood by examining a nozzle with
a constant chamber pressure (Pc) and
applying decreasing ambient back
pressures (P0) (points A→D) on it.
 Again Point A illustrates a limiting case
where P0=Pc, so there is no mass flow
through the nozzle.
 The ambient back pressure (P0) is lowered
to point B and beyond, the static pressure
through the nozzle decreases and the
mass flow through the nozzle increases.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-99
1.5d(ii) Convergent-Divergent Nozzles
 In this range of ambient back pressures, the
flow is fully expanded so Pe=P0. Flow in both
the convergent and divergent portions of the
nozzle is subsonic.
 This trend continues until point C is reached.
At this point the flow at the throat travels at
the speed of sound (M*=1.0).
 Since the flow through the convergent
portion of the nozzle cannot be accelerated
from subsonic velocities to supersonic
velocities; the condi nozzle becomes choked
at all pressure ratios below point C (point D).

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-100
1.5d(ii) Convergent-Divergent Nozzles
 When M*=1.0 at the throat, there are two
possible isentropic solutions for a given area
ratio (A/A*). The flow can either decelerate to
a subsonic exit Mach number or accelerate to
a supersonic exit Mach number.
 Point D (and lower) represents the P0
condition where the flow accelerates to a
supersonic Mach number in the diverging
section of the nozzle. Therefore for P0 lower
than the point D, the pressure will decrease
in both the convergent and divergent sections
of the condi nozzle resulting in supersonic
exhaust flow.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-101
1.5d(ii) Convergent-Divergent Nozzles
 This is the objective of condi nozzle
designs, because a supersonic exhaust
gas velocity greatly increases the thrust of
a propulsion system.
 For P0 in between points C and D, an
isentropic solution is not possible
because shock waves are formed and this
is an irreversible process. In this case,
shock equations would have to be used to
determine the flow properties.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-102
1.5d(ii) Convergent-Divergent Nozzles
 As just stated, condi nozzles are designed to be choked at the
throat (M*=1.0) so exhaust gases can be accelerated to a
supersonic exit velocity in the diverging section.
 The same choke test derived for convergent nozzles can also
be applied to the throat section of condi nozzles.
 Optimal thrust occurs when the nozzle is sized so that the
exhaust gases are fully (or perfectly) expanded (Pe=P0).
 Imperfect nozzle expansion is caused by not having an ideal
nozzle expansion ratio (ε = Ae /A*) for a particular operating
altitude.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-103
1.5d(ii) Convergent-Divergent Nozzles
 The flow is underexpanded if Pe>P0 and overexpanded if Pe<P0
• Underexpansion is caused by a less than optimal nozzle
expansion ratio, resulting in a loss in momentum thrust.
• Overexpansion is caused by having a greater than optimal nozzle
expansion ratio, which may result in flow separation, which form
shocks inside the nozzle. Nozzle performance losses due to
overexpanded flow are generally much larger than losses due to
underexpanded flow.

Aerospace Propulsion Systems Thomas A. Ward Courtesy NASA


© 2010 John Wiley & Sons (Asia) Pte Ltd SR-71B Showing Shock Diamonds in the Exhaust 1-104
1.5d(ii) Convergent-Divergent Nozzles
 Full expansion of an exhaust jet in a fixed-
condi nozzle can only be achieved when it is
operating at its design pressure ratio.
• Consequently, fixed-geometry condi
nozzles are typically only used in missiles
that spend the majority of flight is at a
predictable supersonic cruising velocity. Courtesy National Museum of the USAF
VG Nozzle on the F100-PW-
• Most aerospace propulsion systems with 100 Engine (F-15 Fighter)

condi nozzles are designed with variable


geometry (VG). This allows the area ratio
to be variably optimized over a range of
flight conditions, improving the condi
nozzle’s effectiveness at generating thrust
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-105
1.6 Shock Waves
 A body traveling at M < 1 through a
compressible fluid (such as air)
creates a disturbance that is
propagated throughout the fluid by a
wave traveling at the local velocity of
sound (relative to the body). This
creates gradual changes in the fluid The first panel shows a situation similar to

properties as it approaches the body.


supersonic flow over a body. Like the driver, the
flow has no time to prepare for it. The second case


is like subsonic flow over a body. Like the driver,
However, if the body is traveling at M > the flow ahead of the body gradually changes to
prepare for it.
1 then the fluid is unable to gradually
change ahead of the body. Therefore
the supersonic body induces a sudden
change in fluid properties due to a
shock wave.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-106
1.6 Shock Waves
 Consideration of shock waves is
important in the design of intakes,
nozzles and ducts of aerospace
propulsion systems capable of
supersonic velocities.
 There are two types of shock
waves.
• The simplest type of shock, the
normal shock, occurs normal to
the flow direction.
• An oblique shock occurs at an
inclined angle to the flow
direction.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-107
1.6a Normal Shocks
 Equations have been derived to determine the change in
properties across a normal shock (Appendix C).

Tt1  Tt 2 ~ No change across a shock

2
M 12 
 1
M2 
2
M 12  1
 1

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-108
1.6a Normal Shocks
  1 2   2 
 1  M 
1  M 2
 1 
  1
1
T2  2 

T1 2  2  1 
M 1   
  1 2 

P2 2  M 12    1

P1  1
 1

   1 2   1    1

Pt 2  2
M1   1 
   
Pt1 1    1  2 M 2   1 
M 12 
 2     1 1   1 
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-109
Example 1.4
A normal shock forms on the
intake of an aircraft flying at
Mach 1.6 at 10 km. Assume
 =1.4. Determine the Mach
number (M2), total pressure
(Pt2), static pressure (P2), total
temperature (Tt2), and static
temperature (T2) of air after
the shock.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-110
Example 1.4
Solution
According to the Standard Atmospheric Table, Appendix A,
Table A.1 for 10 km altitude:

P1  26.5 kPa and T1  223.3 K

According to Table C.1, Appendix C (   1.4) for M 1  1.6 :

Pt 2 P2 T2
M 2  0.6684  0.8952  2.820  1.388
Pt1 P1 T1

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-111
Example 1.4
Assuming isentropic flow outside the intake:

 M 12    1   1
Pt1  P1 1  
 2 
1.4

 

 26.5 10 Pa 1 
3 1.6 2
 1.4  1  1.41
 112.6 kPa

 2 

 M 12    1 
Tt1  T1 1  
 2 
 1.6 2  1.4  1 
  223.3 K  1    337.6 K
 2 
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-112
Example 1.4
Therefore:

  112.6 kPa  0.8952   100.8 kPa


Pt 2
Pt 2  Pt1
Pt1

  26.5 kPa  2.82   74.7 kPa


P2
P2  P1
P1

  223.3 K  1.388   309.9 K


T2
T2  T1
T1

Finally, since there is no change in total temperature across


a shock:
Tt 2  Tt1  337.6 K
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-113
1.6b Oblique Shocks
 The methodology of analyzing flow properties across oblique
shocks is very similar to normal shocks. Even though an
oblique shock is inclined at an angle to the flow direction, it
still creates an abrupt change in fluid properties and is
adiabatic.
• Like a normal shock, there is no change in the total
temperature (Tt) across an oblique shock.
• The difference is that an additional variable must be
introduced to account for the oblique shock’s inclination
to the flow direction.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-114
1.6b Oblique Shocks
 It can be shown that there is
no change in tangential velocity
across an oblique shock.

V1t  V2 t

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-115
1.6b Oblique Shocks
 Substituting this into the normal momentum equation yields
equations that show that the components normal to an
oblique shock act just like a normal shock, while the
components tangential to an oblique shock do not change.

 Therefore, the fluid property ratios across an oblique shock


can be determined by calculating the components normal to
the oblique shock and using the normal shock tables

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-116
1.6b Oblique Shocks
 Figures for oblique shocks are shown in Appendix C. Note that
there are two possible solutions (or none at all).
• A strong shock has a large value of θ and a large pressure
ratio across the shock. It generally occurs when the back
pressure of a supersonic flow is extremely high. However, this
system is unstable and will normally degenerate into the
weaker solution. A strong shock will always slow the
supersonic flow velocity to a subsonic speed.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-117
1.6b Oblique Shocks
• A weak shock is one that has a relatively small value of θ, a
smaller pressure ratio across the shock, and a small back
pressure. A weak solution occurs more frequently on aerospace
system designs than a strong shock. Normally weak shocks will
occur on wings, open inlets and planar surfaces. A weak shock
will always slow the flow velocity to a lower but still supersonic
speed.
• This is the type of shock that mostly occurs in propulsion systems.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-118
1.6b Oblique Shocks
• A third possibility is that there is
no solution at all. This can occur
if there is a great enough wedge
angle . In this case the shock
detaches from the body and may
occur in front of it. An example
of a detached bow shock is
shown in the figure.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-119
Example 1.5

Compare the loss in total


pressure ratio incurred by a
two-dimensional, two-shock
spike diffuser and a three-
shock diffuser operating at
Mach 2.0. Assume each
oblique shock turns the flow
through an angle ( ) of 10.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-120
Example 1.5
Solution
(a) Two-shock inlet calculations:
From oblique flow charts (Figures C.1a and b, Appendix C) for
M1=2.0 ,  = 1.4, and d = 10 the weak shock solution is:

  39.4 M 2  1.64
and

Therefore: M 1n  M 1 sin 
  2.0  sin  39.4   1.27

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-121
Example 1.5

The normal shock tables (Table C.1, Appendix C) can now be


used for M1n = 1.27 :
 Pt 2 
   0.9842
 Pt1 

For the normal shock M2 = 1.64 . Then again from the normal
shock tables:
 Pt 3 
   0.8799
 Pt 2 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-122
Example 1.5

So the total pressure recovery across the two-shock inlet is:

 Pt 3 
  0.8779  0.9842   0.864
Pt 3 Pt 2
  
 Pt1  inlet
2 shock Pt 2 Pt1

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-123
Example 1.5
(b) Three-shock inlet calculations:
This is done similar to the one-shock inlet. From the oblique shock tables
(Figures C-1a and b, Appendix C) again for M1=2.0,  = 1.4, and d = 10 :

  39.4 and M 2  1.64

Therefore, once again:

M 1n  M 1 sin   1.27

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-124
Example 1.5

Again using the normal shock tables for M1n = 1.27 :

 Pt 2 
   0.9842
 Pt1 

For the second oblique shock for M2 =1.64,  =1.4, and d = 10 (Figures C.1a
and b, Appendix C),  = 49.4 and M3 =1.28.


M 2n  1.64 sin 49.5  1.25 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-125
Example 1.5

Again using the normal shock tables for M2n = 1.25 :

 Pt 3 
   0.9871
 Pt 2 

For the normal shock, using the normal shock tables for
M3=1.28 :
 Pt 4 
   0.9827
 Pt 3 

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-126
Example 1.5
Therefore:  Pt 4  Pt 4 Pt 3 Pt 2
  
 Pt1  3inlet
shock Pt 3 Pt 2 Pt1

  0.9827  0.9871  0.9842   0.9547

Thus there is about a 10% improvement in total pressure ratio


gained by using the three-shock inlet over a two-shock inlet at
M1=2.0 . If we were to repeat this calculation for M1=4.0, there
would be a 62% improvement. Thus the improvement increases
with higher speeds.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-127
1.6c Conical Shocks
 Supersonic flow about a three-dimensional circular cone is
more complex than a simple two-dimensional wedge, because
after a conical shock the streamlines curve to satisfy the
conservation of mass.
• Therefore a conical shock will be inclined at a lesser angle
to the flow direction than a simple two-dimensional oblique
shock. This means that a two-dimensional wedge will create
a greater flow disturbance than a three-dimensional cone.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-128
1.6c Conical Shocks
• This is because flow cannot pass around the side of a two-
dimensional wedge, since it extends to infinity in the third
dimension.
• Therefore separate flow relations are necessary to analyze a
conical shock shown in Figures C.3, C.4, and C.5 in
Appendix C.

Aerospace Propulsion Systems Thomas A. Ward


© 2010 John Wiley & Sons (Asia) Pte Ltd 1-129
CHAPTER 1 - SUMMARY
 The equations of mass, linear momentum, and energy were
presented and applied to basic engine components.
 Equations for isentropic flow were applied to idealized engine
model components. The effect of Mach number on isentropic
flow was shown.
 Two different types of nozzles were introduced: convergent and
convergent-divergent (condi) nozzles.
• A limiting factor of nozzles is choked flow, which means that
no additional mass can flow through the nozzle.
 Lastly the formation of shock waves in compressible,
supersonic flow was introduced.
Aerospace Propulsion Systems Thomas A. Ward
© 2010 John Wiley & Sons (Asia) Pte Ltd 1-130

You might also like