Lecture Notes Aerospace Propulsion

Download as pdf or txt
Download as pdf or txt
You are on page 1of 104

Lecture Notes on Aerospace Propulsion

Bioengineering and Aerospace Engineering Department


Universidad Carlos III de Madrid
2015-2016

Aerospace Propulsion

Lecture notes

These lecture notes were originally based on a course that has been taught several years at MIT by Manuel
Martnez-S
anchez. Many of the figures have been taken or adapted from the lecture notes from that course and
from the book Aircraft Engines and Gas Turbines by J.L. Kerrebrock published by MIT press.
The lecture notes have been gradually improved upon by Manuel Martnez-Sanchez, Manuel Garca Villalba,
Pablo Fajardo, Mario Merino, and Andrea Ianiro. If you discover any errata, or have any suggestions, please
contact us at:
[email protected]

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Contents
1 Introduction to aerospace propulsion
1.1 Thrust generation and jet propulsion .
1.2 Effect of external expansion on thrust
1.3 Global performance parameters . . . .
1.3.1 Range of aircraft . . . . . . . .
1.3.2 Efficiencies . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

6
. 6
. 7
. 8
. 9
. 11

2 Aircraft Engine Modeling: the Turbojet


2.1 Thrust equation . . . . . . . . . . . . . . . . . .
2.2 Shaft balance for the turbojet . . . . . . . . . . .
2.3 Fuel consumption . . . . . . . . . . . . . . . . . .
2.4 Design parameters. Effect of mass flow on thrust.
2.4.1 Note on Ramjets . . . . . . . . . . . . . .
2.5 Propulsive efficiency . . . . . . . . . . . . . . . .
2.6 Thermal and overall efficiencies . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

13
13
17
17
17
19
20
20

3 Introduction to Component Matching and Off-Design


3.1 Discussion on nozzle choking . . . . . . . . . . . . . . .
3.2 Component matching . . . . . . . . . . . . . . . . . . . .
3.3 Effects of Mach number . . . . . . . . . . . . . . . . . .
3.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5 Compressor-turbine matching. Gas generators. . . . . .

Operation
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

22
22
23
27
28
29

4 Turbofan Engines
4.1 Ideal turbofan model . . . .
4.2 Shaft balance . . . . . . . .
4.3 Velocity matching condition
4.4 Optimal compression ratio .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

31
31
33
33
34

5 Inlets and Nozzles


5.1 Inlets or Diffusers
5.2 Subsonic Inlets .
5.3 Supersonic Inlets
5.4 Exhaust nozzles .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

36
36
36
37
42

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

6 Principles of Compressors and Fans


47
6.1 Eulers equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.2 Velocity triangles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.3 Isentropic efficiency and compressor map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
7 Compressor Blading, design and multi-staging
7.1 Diffusion factor. Stall and surge . . . . . . . . .
7.2 Compressor blading and radial variations. . . .
7.3 Multi-staging and flow area variation . . . . . .
7.4 Mach Number Effects . . . . . . . . . . . . . .
7.5 The Polytropic Efficiency . . . . . . . . . . . .
7.6 Starting and Low-Speed Operation . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Bioengineering and Aerospace Engineering Dept.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

53
53
54
56
57
57
59

v. 2015-2016

Aerospace Propulsion

Lecture notes

8 Turbines. Stage characteristics. Degree of reaction


8.1 Eulers Equation . . . . . . . . . . . . . . . . . . . .
8.2 Degree of Reaction . . . . . . . . . . . . . . . . . . .
8.3 Radial variations . . . . . . . . . . . . . . . . . . . .
8.4 Rotating blade temperature . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

60
60
60
64
64

9 Turbine solidity. Mass flow limits. Internal cooling.


66
9.1 Solidity and aerodynamic loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
9.2 Mass flow per unit of annulus area and blade stress . . . . . . . . . . . . . . . . . . . . . . . . . . 67
9.3 Turbine cooling. General trends and systems. Internal cooling. . . . . . . . . . . . . . . . . . . . 68
10 Film cooling. Thermal stresses. Impingement.
10.1 Film cooling . . . . . . . . . . . . . . . . . . . .
10.2 Impingement cooling . . . . . . . . . . . . . . .
10.3 Thermal stresses . . . . . . . . . . . . . . . . .
10.4 How to design cooled blades . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

72
72
74
74
75

11 Combustion: Combustors and Pollutants


11.1 Combustion process . . . . . . . . . . . .
11.2 Combustor chambers . . . . . . . . . . . .
11.3 Combustor sizing . . . . . . . . . . . . . .
11.4 Afterburners . . . . . . . . . . . . . . . .
11.5 Pollutants: regulations . . . . . . . . . . .
11.6 Mechanisms for pollutant formation . . .
11.7 Upper-Atmospheric Emissions . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

77
77
79
80
81
82
84
86

12 Introduction to engine noise and aeroacoustics


12.1 Noise propagation . . . . . . . . . . . . . . . . .
12.2 Acoustic energy density and power flux . . . . .
12.3 Noise sources and noise modeling . . . . . . . . .
12.4 Jet Noise . . . . . . . . . . . . . . . . . . . . . .
12.5 Turbomachinery noise . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

89
89
90
91
94
95

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

13 Engine rotating structures


97
13.1 Blade loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
13.2 Centrifugal stresses and disc design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
14 Fundamentals of rotordynamics
14.1 Bearings and engine arrangements
14.2 Lumped mass model . . . . . . . .
14.3 Critical speed . . . . . . . . . . . .
14.4 Forces on bearings . . . . . . . . .
14.5 Comments on blade vibrations . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Bioengineering and Aerospace Engineering Dept.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

100
100
101
103
103
104

v. 2015-2016

Aerospace Propulsion

Lecture notes

Main nomenclature

()

General engine and flight magnitudes


Specific impulse [m/s] (NB: also used in [s] in older texts)
Thrust [N]
Drag [N]
Lift [N]
Weight [N]
Fuel mass flow [kg/s]
Mass flow (of air) [kg/s]
Nondimensional mass-flow (with respect to the critical mass flow rate) [-]
Specific fuel comsumption [s/m]
Fuel mass ratio to air mass flow, f = m
f /m
0 [-]
Specific heat value of the fuel [J/kg]
Overall, propulsive and thermal efficiency; = t p [-]
Flight Mach number [-]
Flow Mach number at stage i [-]
Flight velocity [m/s]
Exhaust velocity [m/s]
Aircraft range [m]
Component subindices
Subindex for compressor
Subindex for turbine
Subindex for burner (combustor)
Subindex for afterburner
Subindex for diffuser
Subindex for nozzle
Subindex for fan
Thermodynamic properties: stagnation states and ratios
Stagnation (total) pressure and temperature at station i [Pa, K]
Static pressure and temperature at station i [Pa, K]
Stagnation pressure ratios of compressor, turbine, etc. [-]
Stagnation temperature ratios of compressor, turbine, etc. [-]
Non-dimensional stagnation temperature at station i (normalized with T0 ) [-]
Non-dimensional stagnation pressure at station i (normalized with p0 ) [-]
Non-dimensional stagnation temperature at turbine inlet, t = Tt4 /T0 [-]
turbine inlet to free-stream Stagnation temperature ratio,
p = Tt4 /Tt0 [-]
Sound velocity at station i, [m/s]. For an ideal gas, a = Rg T
Mach number at station i, [-]
+1
 2  2(1)
Choked nondimensional mass flow parameter, () = +1
[-]

u, v, w
V
D
z
R
,

Turbomachinery
Radial, Azimuthal (tangential), and axial velocity components [m/s]
Velocity modulus [m/s]
Diffusion factor [-]
Zweifel coefficient [-]
Degree of reaction [-]
Flow and power coefficients [-]

Isp
F
D
L
W
m
f
m
0; m

m
i
SFC
f
h
, p , t
M0
Mi
u0
u9
R
c
t
b
a
d
n
f
pti , Tti
pi , Ti
c , t , etc
c , t , etc
i
i
t

ai
Mi

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

1
1.1

Lecture notes

Introduction to aerospace propulsion


Thrust generation and jet propulsion

Propulsion
is the1.-Introductory
action of exerting a forward
force F (termed thrust) on a vehicle. According to Newtons
Session
L ecture
second law,
d (mv)
= F + D (v) ,
(1.1)
Propulsion O verview.
dt
this thrust is necessary to accelerate the vehicle (i.e., change its momentum) or to compensate any resistive
ofmovement
exerting in
a forward
force onaauniform
vehicle.velocity
This force
on
forcesPropulsion
that oppose is
tothe
the action
vehicles
order to maintain
(in themust
case react
of aircraft,
this force
is
the
aerodynamic
drag).
something else, like the ground for a ground vehicle. For an aircraft in flight, the reaction is
Newtons third law establishes that the existance of this force is necessarily linked to the existance of a
exerted on air that has been admitted for the purpose, and that is therefore accelerated
reaction force equal in magitude but opposite in direction, acting somewhere else, e.g. the ground for a ground
backwards,
together
with
fuelis products.
On air
a rocket
space,
wherebyno
can be
vehicle.
For an aircraft
in flight,
thesome
reaction
exerted on the
that hasinbeen
admitted
theair
engine
for the
purpose,
and
that
is
therefore
accelerated
downstream,
together
with
some
fuel
products.
On
a
rocket
in
space,
admitted, the reaction can only be exerted on some material that is carried on board and
where there is no air to be admitted, the reaction force can only be exerted on some material that is carried
ex!"##"$%&'%()*(%+!""$,%-+%."%.)##%+""/%)'%)+%012213%'1%"4&#5&'"%'("%25'5&#%6170"%89'(75+':;%
on board
and expelled from the vehicle at high speed. As we will see, it is common to evaluate thrust in terms
of theinmomentum
gained
by the medium
(i.e.,
force),
it must
keptininmind
mind that
that this
this is
terms of the
momentum
gained
bythe
thereaction
medium,
butbut
it must
bebekept
isan
an
indirect procedure, and the thrust force is, by definition, the force exerted by the medium on the vehicle (in
indirect procedure, and the thrust force is truly exerted by the medium on the vehicle (in
complicated ways, to be sure).
complicatedtheways,
to be
sure). propulsion device is the Rocket. In a rocket combustion chamber,
Conceptually,
simplest
aerospace
some propellants carried on board are reacted together to generate heated gas at high pressure, and this gas is
allowed
to flow through
ansimplest
open ductpropulsion
(a nozzle), where
it pushes
the walls
and is pushed
backwards
Conceptually,
the
device
is the forward
Rocket.onSome
propellants
carried
on
by the walls. The speed u9 of the gas as it leaves the nozzle is a measure of the momentum per unit mass it
boardand
are so,
reacted
together
to generate
has gained,
if the mass
flow rate
is m 9 , theheated
thrust gas
is at high pressure, and this gas is allowed to

flow through an open duct 8&%931<<#":;/%.("7"%)'%!5+("+%617.&7$%13%'("%.&##+%&3$%)+%!5+("$%


F m 9 u9 .
(1.2)
backwards by the walls. The speed ue of the gas as it leaves the nozzle is a measure of the
As it can be easily shown by applying the momentum equation to an appropriate control volume traveling with
momentum per unit mass it has gained, and so, if the mass flow rate is !"#$the thrust is
the rocket, this equation is only formally correct if the nozzle is working under pressure-matching conditions.
=>$!"the
%& ' gas
More
gas may
continueoutside
to expand
and accelerate
outside
the engine,
Otherwise,
mayprecisely,
continue tothe
expand
and accelerate
the engine,
and a correction
is necessary,
as
will be
shown
in
section
1.2.
and a correction is necessary, as we will see later.

()'
!"#$%&'
'

Mechanism of thrust generation in a rocket

Figure 1.1: Mechanism of thrust generation in a rocket

In an airbreathing engine, such as a Turbojet, air is admitted at a rate !" ( with a relative
Inspeed
an airbreathing
a Turbojet,
from
thethe
atmosphere
a rate
m
0 rate
withisa
u0 (equalengine,
to the such
flightasspeed),
and air
fuelis isadmitted
added so
that
expelledatmass
flow
relative speed u0 (equal to the flight speed), and fuel is added so that the expelled mass flow rate is m
9 . Then,
!" &expresssion
. Then, aside
from is
the correction for external acceleration, the new expression for thrust
the new
for thrust
F =m
9 u9 m
0 u0 .
(1.3)
is
6

$$$$$) *
%(
Bioengineering
and!"Aerospace
Engineering
Dept.
& %& + !"(

v. 2015-2016

Aerospace Propulsion

Lecture notes

In this equation, we are again assuming that the nozzle is operating at pressure matching conditions. The besic
parts of a Turbojet are the air intake, a compressor, a combustion chamber, a turbine and a nozzle.

#$( )( !

#$3 )3
!
Figure
1.2: Turbojet
schematic
T urbojet
schematic

It is fairly
intuitive
it would
advantageous
against
a larger
air, to
Clearly,
if we move
a largerthat
airflow
m
0 , we be
need
to accelerate it to
lessreact
(i.e., the
exhaust
velocitymass
will beoflower)
to provide
the same
thrust:
this is
more advantageous,
the amount(in
of the
jet momentum
behindlike to
minimize
the
amount
ofintuitively
jet momentum
left behindasuselessly
limit, oneleft
would
uselessly is decreased (in the limit, one would like to react against the whole planet, as in a ground vehicle).
react
the whole
assection
in a ground
vehicle).
This is efficiency
the mainisargument
This idea
willagainst
be understood
moreplanet,
clearly in
1.3.2, where
the propulsive
introduced,for
and
is the replacing
main argument
for
the
by uthe
Turbofan.
In the
turbofan,
the
is split
= replacing
m
coreby
ucore
+
m
bypass
(m
core
+m
air
)u
theFturbojet
theturbojet
Turbofan.
Here,
the
admitted
is split
into coreairair,
which
bypass
bypass
0 admitted
into core air, which reacts with the fuel and generates the torque required to drive the fan, and bypassed air,
reacts with the fuel and generates the torque required for the fan, and bypassed air, moved
moved by the fan. The core may be basically a turbojet with an added low-pressure turbine, and the bypass
byhandles
the fan.
The
coreflow.
mayThe
beturbofans
basically
a turbojet
withinan
added reach
low-pressure
and the
actually
most
of the
that
are nowadays
operation
up to a 9:1turbine,
mass flow
ratio between the bypass and the core streams.
bypass actually handles most of the flow:
The thrust is now (again, aside from external expansion corrections) the sum of the momenta added to both
streams,
F =m
core ucore + m
bypass ubypass m
0 u0
(1.4)
In terms of propulsive efficiency, the next logical step would be to eliminate (to the extent possible) the
core and obtain a simple Propeller, or in a modified form, a Helicopter Rotor. This propeller has to be driven
by an engine, generally the core of a turbojet (optimized to transmit most power to the propeller instead of
increasing the energy of the exhaust gases of the turbojet) or a reciprocating piston engine. The historical order
has been, of course, the reverse (propellers first, followed by turbojets and later turbofans), but this had to do
with details of technological feasibility that you will appreciate better as we progress in this course.

1.2

Effect of external expansion on thrust

pe

We will only discuss in detail the case of a rocket, and the others can be easily understood by extension. We
(pevolume
p0 )Ae
need to calculate the total force exerted by the gas applying the momentum equation to a control
comprising the system. If the pressure p9 at the nozzle exit is not equal to p0 , the undisturbed outside pressure,
the internal gas is pushed forward by the gas outside by a force (p9 p0 )A9 , where A9 is the exit plane area.
This adds a second contribution to the thrust force:

Ae
mu
e

F = mu

+ (p p )A .

9
F = mu
e9+ (p9e p0 0 )A
e.

Bioengineering
and
F =
mc

c Aerospace Engineering Dept.

(1.5)
v. 2015-2016

Schematic of a T urbofan engine


c

ue

The thrust is now (again, aside from external expansion corrections) the sum of the
p e = p0
momenta added to both streams:

p0

It is fairly intuitive that it would be advantageous to react against a larger mass of air, to
minimize the amount of jet momentum left behind uselessly (in the limit, one would like to
react against the whole planet, as in a ground vehicle). This is the main argument for
replacing the turbojet by the Turbofan. Here, the admitted air is split into core air, which
reacts with the fuel and generates the torque required for the fan, and bypassed air, moved
by the
fan. The
core may be basically a turbojet with an added low-pressure turbine,
andnotes
the
Aerospace
Propulsion
Lecture
bypass actually handles most of the flow:

Schematic
of a T urbofan
engine
Figure 1.3: Schematic
of a Turbofan
engine
The thrust is now (again, aside from external expansion corrections) the sum of the
This is sometimes written as F = mc,
momenta
added to both streams: where c is the effective jet velocity. In reflection, one can see that
c really means the speed the gases will achieve after full equilibration of the jet and external pressures, some
distance behind the vehicle. The difference between c and u9 is not very large in general (a few per cent), and it
! at"the
#$%&'(
)%&'(
+#$*+,-..
)*+,-..
#$be
2)3by controling the engine
when/
p9 0#
= p$ %&'(
can
achieved
is literally zero
pressure
matching
condition,
0 ; this1
*+,-..
internal pressure, by variation of the nozzle degree of expansion, or by flying at the appropriate altitude.
For a given jet engine, if the outer (ambient) pressure is fixed and we can vary the nozzle exit area to change
the pressure p9 , maximum thrust is obtained when p9 = p0 . The reason is that if p9 < p0 , the region of the
nozzle interior next to the exit will produce suction (negative thrust), whereas if p9 > p0 , one could add positive
thrust by adding some extra diverging area to the nozzle. This fact will become clear in the study of nozzle
flows. These comments can be extended easily to other engine types.
As a final point, you may be wondering how it is possible to have p9 different from p0 at the exit from the
nozzle. Indeed, this would not be possible if the flow were subsonic there, because there cannot be a significant
pressure difference across a low-speed nearly parallel flow. But if the flow is supersonic, as it almost always is
in propulsive nozzles, the pressure equalization takes place through more complex flow features, such as oblique
shocks anchored at the nozzle lip, and the internal pressure at the exit plane can be different from the outside
pressure.

1.3

Global performance parameters

For all the devices discussed, thrust is proportional to fuel flow rate, and it is advantageous to minimize this
flow rate, or to maximize the ratio known as specific impulse
Is =

Thrust
F
=
Fuel mass flow rate
m
f

(in m/s).

Bioengineering and Aerospace Engineering Dept.

(1.6)

v. 2015-2016

Aerospace Propulsion

Lecture notes

This is the rational definition, but historically, specific impulse has been defined as thrust per unit weight flow
rate of fuel
F
Is =
(in s).
(1.7)
gm
f
In this notes we are going to try to use most of the times the first definition, although students should always
check units carefully to be sure about which definition is being used. For a rocket, the only flow rate is that of
the fuel, but for an airbreathing engine, care must be taken to use the fuel flow rate in the calculation of Is ,
not the air or the total flow rate.
The inverse of Is is also in common use, especially for airbreathing engines, and it is called the Specific Fuel
Consumption (or Thrust Specific Fuel Consumption), often expressed in picturesque units such as pounds of
fuel mass per hour-per pound of thrust force, or the metric equivalent,
SFC or TSFC =

m
f
1
=
Is
F

(in s/m).

(1.8)

EF7$0)G)2GH'1<F)+I6)12
For an aircraft, the simplest measure of performance is the cruise fuel consumption and the resulting range. At
1.3.1

Range of aircraft

one time this was also the critical performance measure for transports, bombers and fighters. This is less true
nowadays for transports because the ranges accessible with modern engines and airframes are in the order of
13,000 km. For bombers, aerial refueling extends the range to the extent that again range is no longer such
a challenge (although in-flight refueling is quite expensive). Range is still important for fighters because the
requirements for high speed and maneuverability conflict with those for long range.

27('+%6J63('+23+6+8J913<+617KJ63'

Figure 1.4: Equilibrium of forces in straight level flight.

E
!
! level flight as shown in Fig. 1.4. The thrust has to compensate the drag,
Consider an aircraft in straight,
F = D, and the lift hasH
to compensate
LMNthe weight, L = W . So that we may write
F =D=D

7:6330@ ;30)1)$)123*+AB5@ CD!!

#$%&'()*+,-./+#'00123*!4

L
L
W
=
=
.
L 560)'7+13+-17(768)+#90)':0+;3)'<76)123+
L/D
L/D

(1.9)
!"

During cruise the rate of change of aircraft weight W is equal to g times the mass flow rate at which fuel is
burned, that is
dW
F
W
= g m
f = g = g
,
(1.10)
dt
Is
Is L/D
9

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

where the first definition of Is , eq. (1.6), has been used.


If we assume that Is and L/D are constant (here we are implying models for both the propulsion system
and the aircraft), then
dt
dW
= g
,
W
Is L/D




W (t)
m(t)
t
log
= log
= g
.
W0
m0
Is L/D

(1.11)
(1.12)

Now we can define the range of an aircraft as R = u0 t, where u0 is the flight cruise speed, and introducing
t = R/u0 into eq. (1.12) we obtain


m0
Is L
log
(1.13)
R = u0
g D
m(R)
where m(R) is the aircraft mass at the range R. This is the Breguet range equation, named after one of
the aviation pioneers, the French aircraft designer and builder Louis Charles Breguet. This equation is also
expressed in terms of the SFC instead of using the specific impulse as


L
m0
1
R = u0
log
.
(1.14)
g SFC D
m(R)
It is useful to divide the initial mass of the aircraft m0 into several parts: empty mass (all the mass required
for the aircraft to operate but the fuel, that is: structure + engines + crew), payload and fuel:
m0 = mempty + mpay + mf uel ,

(1.15)

mempty
mpay
mf uel
+
+
.
m0
m0
m0

(1.16)

or, in terms of mass fractions


1=
If the fuel is all expended at R,

m(R)
mempty
mpay
mf uel
=
+
=1
.
m0
m0
m0
m0

(1.17)

For a fixed structure and engines, we can trade off between payload and fuel, hence between range and payload.
We can now re-write eq.(1.13), as


m0
Is L
R = u0
log
(1.18)
g D
mempty + mpay
and solving for mpay , we obtain


mpay
Rg
mempty
= exp

.
m0
u0 Is L/D
m0

(1.19)

From this we can construct a range vs. payload chart. As an example suppose
mempty = 0.7m0 ,
u0 = 300m/s,
Is = 39200m/s,
L/D = 15,

10

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

so that
u0

Is L
= 1.8 107 m = 1.8 104 km,
g D

(1.20)

and we obtain the chart displayed in Fig. 1.5. Note that this is not a straight line, although it is close. This
graph presents the maximum payload capacity of the aircraft as a function of the range required, showing the
theoretic interchangeability between payload and range. This behavior is obviously limited by the maximum
capacity of the fuel deposits in the considered aircraft.
0.3

0.25

0.2

mpay 0.15
m0
0.1

0.05

0
0

1000

2000

3000

4000

5000

6000

7000

R [km]
Figure 1.5: Range vs. payload chart.

1.3.2

Efficiencies

For airbreathing engines, three kinds of efficiency are in use, that will be discussed more fully later
Overall efficiency, defined as the propulsive power (thrust times flight velocity) over the input thermal
power, which is the product of the fuel flow rate and the fuel heat value (heat of combustion) h, i.e.,
the amount of heat released in combustion per kg of fuel
=

F u0
u0
=
Is ,
hm
f
h

(1.21)

where the last equality follows from the definition of specific impulse.
In the case of the turbojet or the turbofan, the exhaust mass flow can be approximated as
m
9=m
0+m
f 'm
0.

(1.22)

Propulsive efficiency, defined as the ratio of propulsive power (useful propulsion power) to the rate of
addition of jet kinetic energy to the jet
p =

F u0
,
m
0 (u29 u20 ) /2

(1.23)

using here F = m
0 (u9 u0 ), equation (1.23) reduces to
p =
11

2u0
.
u9 + u0

Bioengineering and Aerospace Engineering Dept.

(1.24)
v. 2015-2016

Aerospace Propulsion

Lecture notes

Notice here that p tends to 1 when u9 tends to u0 . For this to mean a non-zero thrust, the mass flow
rate must be large; this is again the argument for the turbofan or the propeller, as opposed to the pure
turbojet: ideally, we would like to exert the reaction force to our thrust in as large an air flow as possible,
to minimize the kinetic energy deposited in the air and maximize p . Observe that for a rocket the
propulsive efficiency is undefined and does not make sense (flight velocity becomes irrelevant, and there
is no external airflow flowing through the engine).
Thermodynamic efficiency, t , defined as the ratio of jet kinetic power to input thermal power. Clearly,
= t p

12

Bioengineering and Aerospace Engineering Dept.

(1.25)

v. 2015-2016

Aerospace Propulsion

Lecture notes

Aircraft Engine Modeling: the Turbojet

All aircraft engines are Heat Engines They use the thermal energy derived from combustion of fossil fuels to
produce heat and then transform it into mechanical energy in the form of kinetic energy of an exhaust jet.
The excess of momentum of the exhaust jet over that of the incoming airflow produces thrust. As such, the
thermodynamic efficiency of the device cannot exceed Carnots efficiency limit,
t 1 Tlow /Thigh ,

(2.1)

here being Thigh the temperature after combustion and Tlow the ambient temperature.
In studying these devices we thus employ two types of models:
Thermodynamic, in which the conversion of thermal energy in mechanical energy from is studied by the
approaches of Thermodynamics. Here the change in thermodynamic state of the air as it passes through
the engine is studied. The detail of the fluid movement within each component is not identified. Rather,
the processes are specified by pressure and temperature ratios. This is the type of model that we derive
in this chapter.
Fluid mechanical, in which we relate the changes in pressure, temperature and velocity of the air, to
the physical characteristics of the engine, e.g. the flow field over the blades of the compressor, in the
combustion chamber and in the turbine.

2.1

Thrust equation

With these ideas in mind, let us first outline a general approach to the modeling of aircraft propulsion systems.
Our general expression for thrust, in which we have a main interest, is
F =m
9 u9 m
0 u0 + (p9 p0 )A9 ,

(2.2)

where m
9 = (1 + f )m
0 includes the fuel mass flow added to the airflow, and f = m
f /m
0 . Subscripts indicate
the standard flow station as shown in Fig. 2.1. We write this more conveniently in dimensionless form as


F
u9
p0 A9 p9
= (1 + f ) 1 +
1 .
(2.3)
m
0 u0
u0
m
0 u0 p0
In our modeling of the aircraft engine we will often assume that the nozzle is operating at pressure matching
conditions (p9 = p0 ), and usually take f  1, so, using the flight Mach number M0 = u0 /a0 , this expression
becomes simply


F
u9
= M0
1 ,
(2.4)
m
0 a0
u0
but it should be recalled that the behavior of the nozzle is somewhat more complex. In practice the deviation from ideal expansion becomes important for supersonic flight. In particular, there can be supersonic
underexpansion, with an exhaust pressure p9 > p0 . This topic will be analyzed in detail in future lectures.
Our tasks in estimating F are then
1. To estimate m
0 , which will depend on the engine front area, flight velocity, etc., and
2. To estimate

u9
. We will start by discussing this first.
u0

Many of the engines we deal with (Turbofans) will have 2 exhaust streams. In this case we apply eq. (2.4)
separately to each stream.

13

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

In the following, we employ the defining relations for the stagnation properties1


1 2
Tt = T 1 +
M ,
2


1 2 1
.
M
pt = p 1 +
2

(2.5)
(2.6)

Remember that a process that does not add or remove heat or work to the fluid always has Tt = const, and
that if Tt is constant and the process is isentropic, then pt is constant too.
Let us begin with a general Turbojet Engine, shown in Fig. 2.1 with the standard station numbering2 .
Station 0 refers to the unperturbed airflow, far upstream.

Figure 2.1: Schematic diagram of turbojet engine. This is the standard station numbering developed by SAE.
And, let us split the engine into a set of Components with functions as follows (in parenthesis, the letter
that will identify the component):
Diffuser (d) or inlet: Brings airflow from the flight Mach number M0 , to the axial Mach number M2 ,
required by the compressor. Ideally S = 0 and Tt = 0, so pt = 0. In reality, especially for M0 > 1,
the process is not isentropic. The diffuser is the region located between flow stations 1 and 2 in Fig. 2.1.
Compressor (c): Raises the pressure (and the temperature) of the airflow by adding mechanical energy to


pt3
Tt3 1
it, as isentropically as possible. If ideal, then
=
, with Tt3 > Tt2 (work added). Losses and
pt2
Tt2
pt3
real effects reduce the achievable
. The compression process in a large engine might be split in several
pt2
steps, e.g. low pressure and high pressure compressor.
Combustor (b): Raises temperature by adding heat (from the chemical energy of the burnt fuel), ideally
at near-constant stagnation pressure (Brayton cycle). In practice, some stagnation pressure losses always
occur. The maximum temperature that can be obtained with the combustion is fixed by the maximum
temperature that the turbine can withstand.
1 Check the companion notes, Basic Relations of Gas Dynamics, for a brief review of stagnation properties and other central
concepts.
2 Students should be aware of slight differences between this station numbering and that used in the book from Kerrebrock.

14

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Turbine (t): Extracts mechanical energy from the airflow, dropping its temperature and pressure, as nearly


pt5
Tt5 1
isentropically as possible. Tt5 < Tt4 (work extracted) with an ideal pressure drop
. The
=
pt4
Tt4
work extracted in the turbine is used to drive the compressor and sometimes also some auxiliary systems.
Afterburner (a): Only present in military turbojets, the afterburner heats air again by burning extra fuel,
at nearly constant stagnation pressure. Here the air temperature can be increased without the limitations
due to the thermal stresses in the turbine.
Nozzle (n): Expands hot gases to produce a high-velocity jet. Station 8 denotes the nozzle throat (section
of minimal area in the nozzle).
In order to calculate the velocity ratio u9 /u0 , we first note that
r
M9 T9
u9
=
.
u0
M0 T0

(2.7)

It is more efficient (and easier) to find the exit Mach number and temperature by keeping track of the
stagnation temperatures and pressures through the several components. The following procedure works for all
aircraft engines, so its worth your paying some attention to the procedure itself, as well as the result.
It is very helpful to define a set of symbols that represent explicitly ratios of the stagnation properties and
distinguish them from the static or thermodynamic properties of the gas, because in general it is the stagnation
properties that most conveniently represent the effect of the components on the fluid as it flows through the
engine. Thus,
A ratio of pt s through a component will be denoted by the symbol and the subindex of that component.
Similarly, a ratio of Tt s will be denoted by the symbol .
A ratio of a stagnation temperature (at one station) to the ambient static temperature T0 will be denoted
by with the corresponding station subscript.
Similarly, a ratio of a stagnation pressure to the ambient static pressure p0 will be denoted by .
So, for the flow upstream of the engine,


Tt0
T0

pt0
p0

1 2
1+
M0
2


1+

1 2
M0
2


= 0 ,

(2.8)

 1

= 0 .

(2.9)

As a particular case, due to its importance in the operation of the turbojet, the turbine-inlet temperature is
represented by
Tt4
= t
(2.10)
T0
or, alternatively, by
=

t
Tt4
Tt4
=

0
Tt0
Tt2

(2.11)

which is more convenient for scaling purposes, since it relates two engine total temperatures, a ratio that is
often independent of ambient conditions. Note that the ideal process in the diffuser is at constant total enthalpy

15

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

(i.e., constant total temperature), since neither work or heat is added to the flow. For the compressor (ideal)
pt3
= c ,
(2.12)
pt2
Tt3
= c ,
(2.13)
Tt2

= c1 ,

(2.14)

= t ,

(2.15)

= t ,

(2.16)

and for the turbine (also ideal by assumption),


pt5
pt4
Tt5
Tt4
t

= t

(2.17)

Now let us use this notation system to develop expressions for the Thrust and Specific Impulse of the
Turbojet Engine. We begin by tracking the changes of stagnation temperature and pressure through the
engine. Temperature accounting:


1 2
Tt9 = T9 1 +
M9 = T0 0 c b t = T0 t t .
(2.18)
2
Pressure accounting:


1 2 1
M9
= p 0 0 c b t
pt9 = p9 1 +
2
From equation (2.19), if p9 = p0 (ideally expanded nozzle) and if b 1, the equation becomes


1
1 2
M9 = (0 c t ) = 0 c t
1+
2


(2.19)

(2.20)

where the second equality assumes that the compression and expansion processes are reversible adiabatics. From
this we find an expression for the exit Mach number,
M92 =

2
(0 c t 1) .
1

(2.21)

It is very important to realize that although this expression for the exit Mach number is written in terms
of temperature ratios, it comes from the pressure changes in the engine (i.e., it was calculated from Eq. (2.19)
alone). This is a general result, namely that the exit Mach number depends on the ratio of jet stagnation
pressure to the ambient pressure, not at all on the temperature ratio T9 /T0 . Now, from eq. (2.18)
T9
t t
t

=
=
=
= b .
1
2
T0

1 + 2 M9
0 c
c

(2.22)

So far these are quite general expressions applicable to any gas stream engine. Substituting eqs. (2.22) and
(2.21) in our expressions for the velocity ratio, eq. (2.7), we have
s
u9
1
2
t
=
(0 c t 1)
.
(2.23)
u0
M0 1
0 c
Finally, the thrust per unit of mass flow (times the speed of sound to make it dimensionless) is
s
F
2
t
=
(0 c t 1)
M0 .
m
0 a0
1
0 c
16

Bioengineering and Aerospace Engineering Dept.

(2.24)

v. 2015-2016

Aerospace Propulsion

2.2

Lecture notes

Shaft balance for the turbojet

So far we have not made this particular to the turbojet engine, because we have not included the relationship
between the compressor and turbine. The fact that distinguishes the turbojet engine from other engines we
may consider later is that the turbine power, m
9 (ht4 ht5 ) equals the compressor power, m
0 (ht3 ht2 ), so:
ht3 ht2 ' ht4 ht5 ,

(2.25)

cp Tt0 (c 1) = cp Tt4 (1 t ).

(2.26)

This equation is known as the power balance in the shaft which can be written as
t = 1

1
0
(c 1) = 1 (c 1).
t

(2.27)

So finally for the Turbojet Engine


F
=
m
0 a0

2.3



2
t
t 0 (c 1)
M0 .
1
0 c

(2.28)

Fuel consumption

We are also interested in the fuel consumption. We get m


f from a combustor heat balance (i.e., the heat power
released by burning m
f fuel per unit time must equal the increase in total enthalpy of the flow across the
combustor):
m
fh = m
9 cp Tt4 m
0 cp Tt3 ' m
0 cp (Tt4 Tt3 ),
(2.29)
where h is the heat released by burning a unit mass of fuel. In other words,
m
f =m
0

cp T0
(t 0 c )
h

so that the fuel-specific impulse, Is = F/m


f is




ha0
1
F
Is =
cp T0 (t 0 c ) m
0 a0

2.4

(2.30)

(2.31)

Design parameters. Effect of mass flow on thrust.

In this section we examine the question of how to choose the key parameters of the engine to obtain some specified
performance at the design conditions, and how the performance varies if these parameters are changed, still
at the design conditions. Later we will look at a complementary question, namely, how the performance of a
particular design changes when conditions are different from design conditions.
With the results that we worked out in the previous section for the Turbojet engine, let us look at the
dependence of F/(m
0 a0 ) on the main parameters, c , M0 and t . We can view them this way
c is a design choice (compressor pressure ratio)
M0 is the flight speed (flight conditions)
t or is the combustor outlet temperature, an operating variable (we can choose how much fuel to burn),
limited by turbine materials to some maximum value.

17

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Assuming again that the nozzle is matched, we can re-write equation (2.28) as
s
 


F
1
2
=
t 1
0 (c 1) M0 .
m
0 a0
1
0 c

(2.32)

we can see that since 0 c > 1, F/(m


0 a0 ) always increases with t . This relationship is displayed in Fig. 2.2 for
a subsonic case. This reflects the fact that, the more energy we add to the flow, the larger the thrust we obtain.
As already indicated, t cannot be increased without limit (turbine survivability sets a maximum value).
3.5

2.5

2
F
m
0 a0 1.5

0.5

0
1

c
F
, vs. compressor temperature ratio c for an ideal turbojet with
m
0 a0
matched exit at M0 = 0.85, and a range of turbine inlet temperatures t = 3, 4, 5, 6, 7, 8. The arrow indicates
increasing values of t . = 1.4.
Figure 2.2: Non-dimensional thrust,

For a given t , what is the variation with c ? By inspection we see that there is a maximum at the maximum
of the bracketed quantity in Eq. (2.32), so at the value of c that satisfies
 


t 1
1

0 = 0
(2.33)
t 1
0 (c 1) =
c
0 c
0 c2
This value is
c |max(F ) =

t
0

(2.34)

This result can be seen to be equivalent to Tt3 = T0 Tt4 , namely, the compressor exhaust should be at the
geometrical mean of the ambient and combustor exhaust temperatures. If it was much lower or much higher,
the T-S diagram of the equivalent Brayton cycle would be too skinny, and enclose too little area (too little
work per unit mass) as shown in Fig. 2.3.
Whether this power is utilized as jet kinetic energy, as in the turbojet, or as shaft power in a turboprop, is
immaterial. Also, as far as this argument goes, the compression Tt3 /T0 can be arbitrarily divided between ram
compression (0 ) and mechanical compression (c ). What is the meaning of this for the turbojet? Putting this
18

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Tt4

Tt4

Tt4

Tt3

Tt3

Tt3
T0

T0

T0

Figure 2.3: Left, too much compression. Middle, optimal compression. Right, too little compression.

value into Eq. (2.32) and the corresponding expression for the specific impulse (2.31), we have the thrust and
Is for engines optimized for thrust per unit of airflow:
s

2
F
2 p
=
t 1 + M02 M0 ,
(2.35)

m
0 a0 max(F )
1

Is |max(F ) =
As an example take: t = 6.25, = 1.4, and

ha0
Cp T0

(t t )

F
m
0 a0


.

ha0
(4.3 107 J/kg)(283m/s)
= 60547m/s
=
gCp T0
(1004J/kgK)(200K)

(2.36)

(2.37)

leading to

q
F
=
11.25 + M02 M0 ,
m
0 a0 max(F )

F
Is |max(F ) = 16157.5m/s
.
m
0 a0 max(F )
2.4.1

(2.38)
(2.39)

Note on Ramjets

Since 0 increases with Mach number, there is an upper limit on M0 reached when the compressor ratiofor
maximum thrust is equal to 1. This theoretical limit is reached for the flight Mach number at which 0 = t .
Considering typical value of t = 9, if 0 > 3 or M0 > 3.9 the preferred engine is a ramjet.
A ramjet is essentially a turbojet without compressor or turbine. In a Ramjet, all compression is due to the
ram (dynamic) effect (0 > 1) so we can put c = 1. No turbine is now needed, so t = 1 as well. The thrust
follows from (2.28)
s
!
r
F
2
t
t
=
(0 1) M0 = M0
1 .
(2.40)
m
0 a0
1
0
0
where eq. (2.8) has been used to simplify the expression. Notice that in our ideal model, both p9 = p0 (matched
nozzle) and pt9 = pt0 (no losses). Taken together, this implies M9 = M0 . But of course, u9 > u0 , because

19

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

T9 > T0 . In fact

Lecture notes

v
u
u
1 2
r
r
r
u
u9
T9
Tt9
t
u Tt9 1 + 2 M0
=
=u
=
=
t Tt0
u0
T0
T

1 2
t0
0
1+
M9
2

(2.41)

which is consistent with eqs. (2.4) and (2.40).

2.5

Propulsive efficiency

The Turbojet engine is attractive for its simplicity and its good thrust behavior at high Mach numbers. Unfortunately it is not very efficient at low Mach numbers, because its jet velocity is too high. To see this, we
consider the Propulsive Efficiency, defined as done in eq. (1.23) (again, for adapted nozzle):
p =

power to airplane
=
power in jet

1
0
2m

2m
0 (u9 u0 ) u0
2u0
F u0
=
=
2 u2 )
2
2
m

(u
u
(u9 u0 )
0
9 + u0
0
9

(2.42)

From this we see that there is a direct conflict between the desire for high jet velocity to give high thrust, and
jet velocity near the flight velocity, to maximize the propulsive efficiency.
In terms of our expression for thrust, since


F
u9
= M0
1 ,
(2.43)
m
0 a0
u0


u9
1
F
F
=
+1=
+ 1,
(2.44)
u0
M0 m
0 a0
m
0 u0
and we can write the expression for the propulsive efficiency in terms of our expression for thrust
p =

2
F
+2
m
0 u0

2
1
M0

F
m
0 a0

(2.45)

!
+2

Since F/(m
0 a0 ) 2 to 3 for low M0 , p is not good for the turbojet at low Mach numbers. We will see later
how this deficiency is remedied by adding a fan to the engine to produce a Turbofan.

2.6

Thermal and overall efficiencies

We have described the evolution of the working fluid in the turbojet as an ideal Bryton cycle: (1) an isentropic
compression, (2) an isobaric (at constant stagnation pressure) heating, (3) an isentropic expansion, and (4)
isobaric (at constant static pressure) heat release (which takes place outside of the engine, as the exhaust
thermalizes with the ambient, closing the cycle). The Thermal Efficiency is defined for the Turbojet Engine as

1
0 u29 u20
m
0 cp (Tt9 T9 Tt0 + T0 )
power in jet
2m
=
=
.
(2.46)
t =
power in fuel flow
m
fh
m
0 cp (Tt4 Tt3 )
Let us write this result in a more convenient way for the considered ideal case. Observe that, according to the
shaft balance, Tt4 Tt9 = Tt3 Tt0 , so we can substitute Tt9 Tt0 = Tt4 Tt3 in the numerator. Hence,
t = 1

20

T9 T0
.
Tt4 Tt3

Bioengineering and Aerospace Engineering Dept.

(2.47)

v. 2015-2016

Aerospace Propulsion

Lecture notes

For the denominator, on the one hand we are considering ideal process in compressor and turbine flow, thus
(1)/
(1)/
Tt4 = T9 (pt4 /p9 )
and Tt3 = T0 (pt3 /p0 )
, and on the other hand, pt3 = pt4 (no pressure drop in
the combustion chamber) and p0 = p9 (pressure matching conditions in the nozzle). Therefore,

t = 1

p0
pt3

 1

=1

T0
T9
=1
.
Tt3
Tt4

(2.48)

Finally we can define an Overall Efficiency as


=

power to airplane
F u0
=
power in fuel flow
m
fh

(2.49)

We see that
= t p .

(2.50)

It is also important that the overall efficiency is directly related to the specific impulse
=

21

F u0
F u0
u0
=
= Is .
m
fh
m
f h
h

Bioengineering and Aerospace Engineering Dept.

(2.51)

v. 2015-2016

Aerospace Propulsion

Lecture notes

Introduction to Component Matching and Off-Design Operation

In last lecture we derived a thermodynamic model that can be used to design a turbojet for a given operating
condition. In this chapter we consider how the engine, already built, behaves in the different operating points.
At this point it is adequate to reflect on which of the many parameters we have introduced (like M2 , c ,
t , t , f , etc.) can be controlled by the pilot, and what are the inter-relationships that determine the others.
This connectivity is in part mechanical, like the shaft power balance (equation 2.27), but it also comes via flow
continuity among components. This topic is usually relegated to the very end of the study of engine components,
where it is introduced under the section of Component Matching. We find it advantageous to move it forward
to this point.
The price to pay for the insight to be gained is the need to introduce one assumption at this point (to
be justified later). This is the assumption that the stators leading to the turbine (the turbine nozzles are
choked). This means the mass flow rate can be written as
pt4 A4
,
m
=m
4 = ()
RTt4
where A4 is the effective flow area of these nozzles and


() =

2
+1

(3.1)

+1
 2(1)

(3.2)

This condition is obtained from the chocked 1D channel flow equation3 .

3.1

Discussion on nozzle choking

The analysis of the flow in nozzles will be done in detail in future lectures. In this subsection an introduction
to the operation of nozzles in chocked conditions is presented. If we want to have a supersonic adapted exhaust
(M9 1), from eq. (2.21) we must have,
2
(0 c t 1) 1,
1

(3.3)

which may not be satisfied at low power and/or low Mach number.
Equation (3.3) is the condition for the exhaust to be supersonic, with4 p9 p0 . It involves both the
compressor and the turbine temperature ratios, but we can eliminate the turbine ratio using the shaft balance
(Eq. 2.27), so that the condition is now


c 1
+1
0 c 1

.
(3.4)

2
which, in case of equality makes M9 = 1 while still p9 = p0 . This limit can be rearranged into a quadratic
equation for c


+1
c2 ( + 1)c +
=0
(3.5)
2
0
with the two solutions
s
2 

+1
+1
+1
+,

.
(3.6)
c
=
2
2
2
0
It can be verified that c must be in the range between these two roots to ensure M9 > 1. The c+ is normally
very high, so the relevant condition is c . Values of c are tabulated in Table 3.1 as a function of the flight
Mach number and of the parameter .
3 See

the additional materials on basic gas dynamics relations for a review of these equations
the following lectures we will see that the only possible situation for having a supersonic jet at the exhaust of a nozzle is to
have p9 > p0 . The use of a convergent-divergent nozzle is required.
4 In

22

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

=4
1.296
1.066

M0 = 0
M0 = 0.85

=6
1.253
1.059

=8
1.234
1.056

Table 3.1: Values of c


These values are fairly low compressor ratios, even for stationary (M0 = 0) engine conditions, so the
assumption of a choked nozzle is a good one in general. Whether or not the nozzle is also matched is a different
question, as noted before.

3.2

Component matching

Passing the same flow through two choked apertures (the turbine inlet, 4, and the nozzle throat, 8) in series
imposes very strong constraints on the flow conditions. The flow can be expressed at the throat as
pt8 A8
m
=m
8 = ()
,
RTt8

(3.7)

and equating equations (3.1) and (3.7)


pt8
pt4

Tt4
A4
=
.
Tt8
A8

(3.8)

For a non-afterburning turbojet, pt8 ' pt5 and Tt8 = Tt5 , therefore
A4
t
,
=
t
A8

(3.9)

and if the turbine is ideal, t = t1 , and we obtain



t =

A4
A8

 2(1)
+1
(3.10)

and then

t =

A4
A8

2
 +1

(3.11)

This is a strong result: as long as both the turbine nozzles and the exhaust throat remain choked, the turbine
maintains the same pressure and temperature ratios (same operating point), regardless of fuel flow, Mach
number, altitude, etc. We can now trace the variability of other quantities:
1. Compressor ratios. In terms of = Tt4 /Tt0 , equation (2.27)
c
c

1 + (1 t ) ,

= c

(3.12)
(3.13)

Thus c and c do vary, but only as a function of the single quantity , c = c () for a given engine.
2. Dimensionless air flow. The flow at compressor inlet is generally subsonic, so we express the flow rate
there as
pt2 A2
m
=m
2 = ()
m2 (M2 ),
(3.14)
RTt2
23

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

with

+1
2(1)
+1

m2 (M2 ) = M2
.

1 2
1+
M2
2

(3.15)

The dimensionless flow function m2 (M2 ) (mass flow rate as a fraction of the critical mass flow, m
2 )
increases to a maximum of 1 when M2 = 1, then decreases again, as shown (for = 1.4) in Fig. 3.1.

0.8

0.6

m2
0.4

0.2

0
0

M2
Figure 3.1: Dimensionless function m2 as a function of M2 for = 1.4.
Equating (3.14) to (3.1), we see that
pt4
m2 =
pt2

Tt2 A4
.
Tt4 A2

(3.16)

For an ideal combustor, pt4 = pt3 , and so, using c = c1 , Tt2 = Tt0 ,

c1 A4
m2 =
.
A2

(3.17)

Since c = c (), we see now that m2 = m2 () as well. This is very useful for scaling from one operating
condition to another.
3. Mach number at compressor inlet (M2 ). Returning to equation (3.14), we see that a certain dimensionless
mass flow, for a given gas, corresponds to two possible Mach numbers (the subonic and the supersonic
solution). Since the non-dimensional mass flow is only function of , M2 = M2 () (the supersonic solution
for M2 can be disregarded).
4. Fuel/air ratio. The combustor heat balance is
m
fh = m
0 Cp (Tt4 Tt3 ) = m
0 Cp Tt2 ( c )
24

Bioengineering and Aerospace Engineering Dept.

(3.18)

v. 2015-2016

Aerospace Propulsion

Lecture notes

and using Tt2 = Tt0 and f = m


f /m
0,
fh
= c ()
Cp Tt0

(3.19)

so this quantity is another function of alone. But notice that, for a given fuel (h)and gas (Cp ) and at
a fixed T0 , f itself does depend on M0 .
5. Throat pressure (normalized). With the hypothesis considered in this section (ideal turbojet)
p8
p8 pt8 pt5 pt4 pt3 pt2
=
,
pt0
pt8 pt5 pt4 pt3 pt2 pt0

(3.20)

is reduced to

p8
p8 pt5 pt3
=
.
pt0
pt8 pt4 pt2
Introducing the chocked flow condition at the nozzle throat
p8
pt8
pt3
pt2
pt5
pt4

(3.21)

! 1
1
1+
1
2

2
+1

 1

(3.22)

= c = c1

(3.23)

= t = t1

(3.24)
(3.25)

we obtain
p8
=
pt0

2
t c ()
+1

 1

(3.26)

which is yet another function of alone.


6. Thrust (matched nozzle). We already have equation (2.28), but it is sometimes better to normalize thrust
by the total free-stream pressure on the compressor inlet, pt0 A2 , which is known from flight conditions.
If p9 = p0 (e.g. variable area nozzle, or just design point for a fixed nozzle),
2

F
m(u
9 u0 )
pt0 A2 (u9 u0 )
=
= m2
.
pt0 A2
pt0 A2
RTt0 pt0 A2

We already had an expression for u9 , from eq. (2.23)


s
2

(0 c t 1) .
u9 = a 0
1
c

(3.27)

(3.28)

and using

a0
=
RTt0

we obtain
r
2 = m 2

"s

RT0
=
RTt0

#
2 (0 c t 1)
M0 .
1
c

(3.29)

(3.30)

Here the quantities m2 and c depend on only, but we can see that the Mach number M0 appears
1 2
explicitly (as M0 and as 0 = 1 +
M0 ), so the normalized thrust 2 depends on both and M0 .
2
25

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

7. Thrust (convergent-only, underexpanded nozzle). We now have M9 = M8 = 1, but p9 = p8 > p0 so




F
(u9 u0 )
p9
m(u
9 u0 ) + (p9 p0 )A9
p0 A9
2 =
+
=
= m2

,
(3.31)
pt0 A2
pt0 A2
pt0
pt0 A2
RTt0
and this time M9 = 1, leading to
u9 =

r
RT9 = R

2
Tt5 =
+1

2
RTt0 t
+1

(3.32)

u9
depends on alone. Since we also know that m2 and p9 /pt0 are functions of alone, it
RTt0
makes sense to separate out equation (3.31) in the form

 

u9
p 9 A9
u0
p 0 A9
2 = m2
+
m2
+
,
(3.33)
pt0 A2
pt0 A2
RTt0
RTt0
"
# "
#

r

 1
r
2
A9

2
1 A9
2 = m2
m2 M0
.
(3.34)
t +
t c
+
+1
+1
A2
0
1 A2
so that

In equation (3.34) the first bracket is a function of only, 2 (). Once again, the normalized thrust
depends on both, and M0 , but the structure is fairly simple, and in particular, the portion 2 of 2
(neglecting the incoming momentum and the external pressure) is a function of alone. This portion can
be very easily scaled between conditions, and the rest can be subtracted separately.
8. The Operating Line in the compressor map. Compressor performance is typically presented as a map of
c vs. m2 , with lines of constant normalized rotational speed and c , the isentropic efficiency of the
compressor (to be defined in future lectures), superimposed. The details are the subject of later Lectures,
but the general shape is as shown in Fig. 3.2, where the flow and speed variables are renormalized by the
Design values.
Actually the nominal operating line shown in the figure is not a property of the compressor, but rather
of the rest of the engine. We can calculate this line with the information we have now, before deciding
what particular compressor to use. From eq. (3.17),

A4 c1
,
m2 =
A2

(3.35)

and from the shaft power balance (Eq. 3.12),


c 1
,
1 t
where we recall that t is fixed for a fixed geometry. Eliminating ,
r

A4 1
1 t
m2 =
c
,
A2
c 1
=

(3.36)

(3.37)

or, in terms of c

A4
m2 =
c
A2

1 t
1

(3.38)

c 1
which is the equation for the operating line (written in reverse).
If the compressor is already available, we see from eq. (3.38) that we can adjust the nozzle area A4 to
place this line in a good place on the map, i.e., below the stall line and through the best efficiency
points. Since m2 depends on = Tt4 /Tt0 , varying Tt4 moves the operating point along the operating line,
and this is what the pilot does with the throttle stick to power the engine up or down. At each selected
, the engine settles to a c , a M2 , a (normalized) rotation rate, etc.
26

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure
The corrected airflow
p 3.2: Performance map for a typical high-pressure-ratio
p compressor.
m
0 T2 /298.15K/(p2 /101325Pa) and corrected rotational velocity N T2 /298.15K are customarily used in these
diagrams to account for ambient condition variations at the entrance of the compressor.

3.3

Effects of Mach number

If we look at operation of a given engine at different flight Mach numbers, we may try to maintain the
same non-dimensional conditions throughout, which, as we have seen, can be done by maintaining for
example a constant compressor inlet Mach number M2 . This, in turn guarantees a constant = Tt4 /Tt0 ,
but since now we have a varying Mach number, so that Tt0 increases with M0 , we may find that the
turbine inlet temperature Tt4 needs to become too high at the higher Mach numbers. For example, Tt4
would have to be 1.8 times higher at M0 = 2 than at static conditions, and 2.25 times at M0 = 2.5.
A more reasonable assumption is that the ratio t = Tt4 /T0 can be maintained the same at all Mach
numbers, since at least in the tropopause (between 11 km and 20 km above mean sea level), T0 is almost
t (1 t )
invariant. The compressor temperature ratio now follows from c = 1 +
, where the numerator
0
is a constant; thus, c will be lowered as the Mach number increases,
 less strongly than as would
but
be required in order to maintain maximum thrust per unit flow
t /0 . The flow parameter is now
determined by Eq. (3.17), i.e. compressor-turbine flow matching, and then the compressor-inlet Mach
number from Eq. (3.15). Once these parameters are known, we can use Eq. (3.27) to calculate the
normalized thrust; since we are interested in the effect of Mach number, it makes sense to re-normalize

27

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

thrust by p0 A2 , or

3.4

Lecture notes

F
= 2 01 .
p0 A2

(3.39)

Examples

A note on : the near-constancy of the engine operating point


In this section different examples of aircraft operation are given in order to show the possibility of maintaining a near-constant value of the parameter . Two important points in the flight envelope of an
aircraft engine are:
(a) Take-off conditions (M0 ' 0.25, T0 ' 290K) , and
(b) End-of-climb conditions (M0 ' 0.85, T0 ' 220K).


Using = 1.4, the total temperatures are Tt0 = 290 1 + 0.2 0.252 = 294K (take-off) and Tt0 =
220 1 + 0.2 0.852 = 252K (end of climb). Suppose the engine is dimensioned for end-of-climb, which is
common, and that the peak temperature Tt4 , which will have to be maintained for many hours of cruise,
is selected at a conservative Tt4 = 1600K. We then have = Tt4 /Tt0 = 6.35 at this condition. If we now
decided to maintain = 6.35 also for take-off, we would need then Tt4 = 6.35 294 = 1868K. While this
is too high for long-term operation (creep, corrosion), it may be acceptable for the few minutes per cycle
that the engine will be at take-off maximum power. If this actually done, the engine operates at a fixed
nondimensional point all the way from take-off through start of cruise.

Consider now a commercial jet in a long cruise. As the fuel is consumed and the weight decreases,
1
the lift must decrease L = 0 u20 Aw cL . Now, the lift coefficient will be kept close to that for optimum
2
aerodynamic efficiency, L/D|max , and the Mach number M0 is unlikely to change much, as it will stay just
below the transonic drag peak, and so u20 will be proportional to T0 due to the speed of sound variation.
Together with the density part of lift, we can see that the ambient pressure p0 must be decreasing in
proportion to the airplanes weight, i.e., the plane must be climbing gradually. Turning now to the forward
force balance, given a constant L/D, the drag, and hence the engine thrust, must also be decreasing in
time in the same proportion as the ambient pressure. Therefore, from Eq. (3.27), the nondimensional
thrust 2 (, M0 ) will remain constant, and since M0 does too, the peak temperature ratio will also
remain constant, and with it all the important ratios like c , M2 , etc.
In other words, may not vary much among (important) flight conditions, and the engine will be operating
at a fixed nondimensional condition (constant compression ratio, nondimensional flow, compressor inlet
Mach number, etc.). But of course, the dimensional quantities (flow rate, peak pressure, etc.) will be
different, depending on p0 , etc.
A numerical example
We want to analyze the parameters during the engine operation at different regimes (different values of
M0 ), but keeping a constant t = 7, or Tt4 = 1540K (T0 = 220K) in the stratosphere. The geometry of
the engine must have been specified in advance. This means that the turbine temperature ratio (Eq. 3.10)
is a known fixed number. For the example, we select t such as to obtain maximum thrust at M0 = 1.
From the shaft balance equation,
0
t = 1 (c 1)
(3.40)
t

and we put now 0 = 1.2 and c = 7/1.2 = 2.2048 (at M0 = 1). This fixes t = 0.7935. Similarly, the
area ratio A4 /A2 must have been fixed, and we select it here so as to obtain at M0 = 1 a compressor-face
Mach number M2 = 0.5, which, from Eq. (3.15) implies m2 = 0.7464.

28

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

In order to analyze how m2 is scaled at different conditions, using eq. (3.17) we can write,
 3.5 r
c
0
m2 = 0.7464
2.2048
1.2

(3.41)

and the rest of the steps are as described above. The table 3.2 summarizes the results.
M0
0
c
m2
M2
2
F/(p0 A2 )

0
1
2.4458
0.9796
0.8486
2.9117
2.9117

1
1.2
2.2048
0.7464
0.5
1.5531
2.9399

2
1.8
1.8032
0.4523
0.2737
0.5795
4.534

2.5
2.25
1.6426
0.3648
0.2172
0.3503
5.985

Table 3.2: Summary of results of the numerical example


We find that at a fixed altitude the thrust is nearly constant up to Mach 1, then it increases rapidly.
Actually the increase is less rapid than this simple model predicts, because of losses in the supersonic flow
in the engine inlet. The previous model has been computed assuming constant altitude. However, we
should note that an aircraft normally flies at increasing altitude as the Mach Number increases, so that
dynamic pressure p0 M02 is roughly constant. In this case the change in F between Mach 1 and Mach 2
is actually a thrust reduction.

3.5

Compressor-turbine matching. Gas generators.

Assembling the compressor and the turbine with a combustor between gives us a gas generator, which is the heart
of any gas turbine engine. Once we understand its behavior we can appreciate most of the real characteristics
of aircraft engines, and graduate from thinking of them as a lot of abstract equations.
Actually, we have already done in the previous section most of the work needed to understand the performance of an ideal Gas Generator, including the important concept of the Compressor operating line, which
is set by the flow passing characteristics of the rest of the engine. We summarize here the main findings
1. If the nozzle and the turbine stators are both choked, the turbine temperature ratio is fixed once the flow
area ratio A4 /A8 is set.
2. One additional single parameter is sufficient to specify all the other gas generator parameters. This can
be the temperature ratio = Tt4 /Tt2 , or the compressor temperature ratio c , or the engine-face Mach
number M2 , or the normalized air mass flow, or the fuel flow ratio.
3. A Compressor Operating Line in the plane of compressor pressure ratio vs. normalized flow can be
calculated once the ratios of all the flow areas are known. For a given choice of one of the parameters
listed above, a point is selected along this operating line.
4. All the above is independent of the compressor specifics. After the compressor has been selected, its performance map (pressure ratio vs. normalized flow) contains normalized rotational speed lines as additional
information. This parameter is therefore to be added to our list of possible parameters (as is done in the
figure at the end of this lecture), each of which uniquely specifies the state of the gas generator.
It follows from the above that a single degree of freedom is left to the pilot (or to the engine controller),
unless geometry can be varied. It is probably most intuitive to think of this unique freedom as the normalized
fuel factor, or the peak temperature ratio Tt4 /Tt2 , since these closely relate to the engine throttle control.
29

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

It is to be noted, however, that many idealizations have been made to obtain these simple results. If the
turbine or the nozzle un-choke, or if the engine inefficiencies are rigorously accounted for, the overall detailed
behavior is more complex, but its main qualitative features are not too different.
Using these ideas, one can generate and plot a set of Gas Generator Characteristics, such as those below.
With this Gas Generator, we can analyze several engines
1. Turbojet
2. Turbofan
3. Turboprop
4. Unducted fan
5. Helicopter-Turboshaft
Notice that the single free variable chosen for this particular plot is the normalized rotational speed (as a
fraction of its design value). The quantity in the denominator is the non-dimensional compressor-face temperature 2 = T2 /T0 , because the blade speed r is made non-dimensional with the speed of sound at station
2.

Figure 3.3: Pumping characteristics for a gas generator with Tt4 /Tt2 = 6

30

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Turbofan Engines

In 2.4 we saw that for low M0 , p is not very high for turbojets. In essence, there is too much kinetic energy in
the exhaust jet (per unit mass). This is the main reason for using the Turbofan engine. The turbofan separates
the inflow into a core airstream and a secondary or bypass airstream. The core stream goes into a gas generator
which is essentially a turbojet. The bypass stream is simply moved by a fan, which is essentially a big ducted
propeller that increases the total pressure and temperature of the air. This fan is driven by the turbine inside
the core part of the engine, which usually has two shafts: the inner shaft rigidly connects the low-pressure
turbine with the fan, and the outer shaft (or spool) connects the high-pressure turbine with the high-pressure
compressor. The spool and the shaft rotate freely using a bearing system. The jets that result from the core and
bypass streams can be merged at the exit or before a possible afterburner stage, or be released independently
into the ambient.
If it is designed for subsonic cruise flight it looks like the sketch of Fig. 4.1.

13

19

Figure 4.1: Schematic diagram of turbofan engine


If designed for both subsonic cruise and for supersonic flight with afterburning it looks more like Fig. 4.2.
In both of these diagrams the inlet has been greatly simplified, of course. The numbering of the stations
follows the standard SAE AS755: numbers in the bypass stream are prepended with a 1, to express that this
is a different air stream from the core one (could be regarded as prepended by index 0).

4.1

Ideal turbofan model

Let us now see how we can model these engines thermodynamically, following the same process as with the
turbojet. The total thrust of the device is given by Eq. 1.4. Notice first that we now have two air streams: the
core stream through the gas generator, with a mass flow m,
and the bypass stream, with a mass flow that can
be written as a fraction of the former, m.
We call the bypass ratio (BPR). To obtain a simplified model of
the device sketched in Fig. 4.1, we will take the following assumptions, which will be familiar from the turbojet
model:
isentropic processes in the fan, compressor, turbine
31

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

13

Figure 4.2: Schematic diagram of turbofan engine with afterburner


constant total pressure through combustor
fuel mass flow neglected with respect to air flow
the nozzle of the core jet and the bypass stream are both adapted to the ambient pressure (p9 = p19 = p0 ).
Of course, this may be not the case in many circumstances: as in the turbojet, a fixed convergent nozzle
is more likely to be sonic and under-expanded, namely, to have M8 = M18 = 1 and p9 , p19 > p0 . The
derivation of the model with non-matched conditions is left to the student (like in the turbojet case).
Let us start with the core jet. With these hypotheses, the expression for the thrust contribution of the core
jet is the same as for the turbojet before applying the shaft balance. Thus from equation (2.24) we get for the
core jet5
s
Fc
2
t
(0 c t 1)
=
M0 .
(4.1)
ma
0
1
0 c
For the bypass stream, there is no turbine, so let us repeat the argument


1 2
M19 = T0 0 f ,
Tt19 = T19 1 +
2



1 2 1
pt19 = p19 1 +
M19
= p0 0 f = p0 (0 f ) 1 .
2
So, for p19 = p0

1 2
M19 = 0 f ,
2
r
2
=
(0 f 1).
1

1+
M19

Using (4.4), from eq. (4.2) we also have that T19 = T0 so



 r
u19
2
FBP
= M0
1 =
(0 f 1) M0 .
ma
0
u0
1

(4.2)

(4.3)

(4.4)
(4.5)

(4.6)

5 Quick observation: = T
t13 /Tt2 . For congruency with the definitions of previous chapters, we keep c = Tt3 /Tt2 , including
f
in it the possible compression of the core stream due to the central part of the fan.

32

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Adding this thrust to the thrust of the core jet, we find the total thrust
"s
#

r
F
2
2
t
=
(0 c t 1)
M0 +
(0 f 1) M0
ma
0
1
0 c
1

4.2

(4.7)

Shaft balance

Now we need t , which has to be calculated with the applicable shaft balance in this case. We will see later
that most engines have two shafts, so each will set a separate balance equation. For now, however, we will lump
their power together; at off-design conditions, this would have to be modified. Applying the energy equation to
the compressor, fan and turbine, and forcing the work extracted by the latter to be consumed by the former,
mC
p (Tt4 Tt5 )

= mC
p (Tt3 Tt2 ) + mC
p (Tt13 Tt2 ),

t (1 t )

(4.8)

= 0 (c 1) + 0 (f 1),

(4.9)

0
[(c 1) + (f 1)].
t

(4.10)

leading to
t = 1

Substituting this gives us our result for the thrust of the turbofan. It doesnt really help to carry out the
substitution at this point. Instead, let us think about how to simplify the expressions to make them more easily
understandable.

4.3

Velocity matching condition

For this engine there are more parameters than for the turbojet
t , c - as before
, f - characterizing the fan flow.
We can relate some of the parameters to 
others by noting
 that the highest propulsive efficiency is realized
momentum
when u19 = u9 : this maximizes the ratio of
in the jets6 . Setting the two velocities to be the
energy
same is referred to as velocity matching, and is usually imposed as a design constrain for the nominal operation

point. For this situation, the two


s in the thrust equation are equal, and this requires that:
(0 c t 1)
6A

t
= 0 f 1
0 c

(4.11)

simple derivation: for two streams 1 and 2,


momentum

kinetic energy

m
1 u1 + m
2 u2 ,
1
1
2
m
1 u1 + m
2 u22 .
2
2

Now we maximize the momentum at a constrained energy. Using a Lagrange multiplier , we form the auxiliary function


1
1
=m
1 u1 + m
2 u2 +
m
1 u21 + m
2 u22 ,
2
2
and equate to zero the derivatives w.r.t. u1 and u2 , leading to
m
1 + m
1 u1

0,

m
2 + m
2 u2

0.

1
So that u1 = u2 =
.

33

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Solving for f as a function of , using eq. (4.10), after some algebra we obtain

f =

1 + t + 0 (1 + c )
0 (1 + )

t
0 c

(4.12)

When this equation is satisfied, the thrust equation becomes simply


r

2
F
= (1 + )
(0 f 1) M0
ma
0
1

4.4

(4.13)

Optimal compression ratio

Now, as for the turbojet, there is still the choice of the compression ratio. Generally we want to choose it for
maximum power, which in this case for given turbine inlet temperature t and bypass ratio means maximum
thrust. To maximize F in this velocity-matching condition, we choose c to maximize f , since F increases
monotonically with f . From the expression for f , eq. (4.12),

t
t
f
= 0 c |max(F ) =
= 0 +
.
(4.14)
2
c
0 c
0
Notice that this is precisely the same result as for the Turbojet. Substituting in the expression for f
f |max(F )

F
m
0 a0 max(F )

2
t 1
+1
=
0 (1 + )
v
u
u 2
= (1 + ) t
1

(4.15)

!
2
t 1
+ 0 1 M 0
1+

(4.16)

Now we have 3 parameters,


t - which we set at the maximum feasible value
M0 - flight speed
- the prime variable distinguishing the turbofan
As before, for u9 = u19 we still have:
propulsive =

2
u9
+1
u0

2
F
ma
0
+2
M0 (1 + )

(4.17)

The variation of F/(ma


0 ) and p with M0 and is shown in Figures 4.3 and 4.4 for t = 6.25. Of course for
the higher bypass ratios we are really only interested in the range of M0 < 1, but the lower bypass and higher
M0 range may be of interest for a supersonic transport.

34

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

10
9
8

20

7
6
5

F 5
ma
0

4
3
2
1
0
0

0.5

1.5

2.5

M0
Figure 4.3: Variation of F/(ma
0 ) with M0 and for t = 6.25 and = 1.4.

1
0.9

0.8
20
0.7
5
0.6
1

0.5
0.4
0.3
0.2
0.1
0
0

0.5

1.5

2.5

M0
Figure 4.4: Variation of p with M0 and for t = 6.25 and = 1.4.

35

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Inlets and Nozzles

5.1

Inlets or Diffusers

While the Gas Generator, composed of the compressor, combustor and turbine, is the heart of any gas turbine
engine, the overall performance of the propulsion system is strongly influenced by the inlet and the nozzle. This
is especially true for high M0 flight, when a major portion of the overall temperature and pressure rise of the
cycle are in the inlet, and a correspondingly large part of the expansion in the nozzle. So it is important to
understand how these components function and how they limit the performance of the propulsion system.
Inlets or Diffusers
These two titles are used interchangeably for the component that captures the oncoming propulsive streamtube and conditions it for entrance to the engine. The function of the inlet is to adjust the flow from the ambient
flight condition, to that required for entry into the fan or compressor of the engine. It must do this over the
full flight speed range, from static (takeoff) to the highest M0 the vehicle can attain. Comparing the simple
diagrams of subsonic and supersonic inlets, we can appreciate that the subsonic inlet has the simpler task:

Figure 5.1: Schematic diagram of diffusers


For any inlet the requirements are two
To bring the inlet flow to the engine with the highest possible stagnation pressure. This is measured by
t2
the Inlet Pressure Recovery, d = ppt0
.
To provide the required engine mass flow. As we shall see the mass flow can be limited by choking of the
inlet.

5.2

Subsonic Inlets

The subsonic inlet must satisfy two basic requirements:


Diffusion of the free-stream flow to the compressor inlet condition at cruise.
Acceleration of static air to the compressor inlet condition at takeoff.
There is a compromise to be made because a relatively thin lip aligned with the entering flow is best for
the cruise condition. This is to avoid accelerating the flow, already at a Mach number of the order of 0.8, to
supersonic speeds that will lead to shock losses. But a more rounded lip will better avoid separation for the
takeoff condition because the air must be captured from a wide range of angles:
Usually subsonic diffusers consist of a divergent duct. Therefore, the minimum area is located at the inlet.
The inlet area, A1 , is set to just avoid choking (M1 < 1) when M2 (compressor inlet) has its largest value (this
36

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 5.2: Schematic diagram of a subsonic diffuser


is set by the maximum blade Mach number and the blade shape). Since the total temperature and pressure are
common, we must have
m1 (M1 )A1 = m2 (M2 )A2 .
(5.1)
A well designed subsonic inlet will produce a stagnation pressure recovery d in the order of 0.97 at its
design condition.

5.3

Supersonic Inlets

At supersonic flight speeds the pressure and temperature rise in the inlet can be quite large. For the best
thermodynamic efficiency it is important that this compression is as nearly reversible (isentropic) as possible.
At a flight Mach number of 3 the ideal pressure ratio is
pt0
0 =
=
p0
while the temperature rise is
0 =

1 2
1+
M0
2

 1

= 36.7

Tt0
1 2
=1+
M0 = 2.8
T0
2

(5.2)

(5.3)

For the turbojet cycle the compression is partly in the inlet and partly in the compressor:
If the diffusion from point 0 to point 2 is not reversible, the entropy increases, and this results in a lower
value of pt2 . This has two effects:
The expansion ratio of the nozzle is decreased, so the jet velocity and thrust are lower.
The lower pressure at point 2 limits the mass flow through the compressor to a lower value than for
isentropic diffusion.

37

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

There is, thus, a double penalty for losses in the inlet. Unfortunately some losses are inevitable and are higher
as M0 is larger. To see why, we will discuss the flow in supersonic diffusers, beginning with the simplest.
1. Normal-Shock diffuser
All existing compressors and fans require subsonic flow at their inlet with 0.5 < M2 < 0.8 at high power
conditions. So the inlet must reduce the flow Mach number from M0 > 1 to M2 < 1. The simplest way
to do this is with a Normal Shock Wave.

Figure 5.3: Normal shock wave


Here M1 < 1 is entirely determined by M0 , according to the normal shock relation

M12

1 2
M0
2
=
1
M02
2
1+

(5.4)

The stagnation and static pressures are also determined by M0


pt0
pt1
p1
p0

 1 

1 1 ( 1)M02 + 2 1
2M02

+1
+1
( + 1)M02

2
M02 1
1+
+1


=
=

(5.5)
(5.6)
(5.7)

For low supersonic speeds, such diffusers are adequate because the stagnation pressure loss is small, but
at M0 = 2, pt2 /pt0 .71, a serious penalty, and at M0 = 3, pt2 /pt0 .32. For example the F-16 fighter
has a simple normal shock diffuser, while the F-15 has an oblique shock diffuser such as will be discussed
next.
2. Oblique - Shock diffusers
The losses can be greatly reduced by decelerating the flow through one or more oblique shock waves,
the deflection and the pressure rise of each being small enough to be in the range where the stagnation
pressure ratio is close to unity. It is very important to understand that an Oblique Shock Wave is in fact
just a normal shock wave standing at an angle with respect to the flow. For an oblique shock wave the
relevant Mach number is that corresponding to the normal component of the velocity.
38

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

1
0.9
0.8
0.7
0.6
0.5

!
!
!
"#$!%#&!'()*$'#+,-!')**.'/!'(-0!.,11('*$'!2$*!2.*3(24*!5*-2('*!40*!'426+24,#+!
0.4
)$*''($*!%#''!,'!'72%%/!5(4!24!8#!9!:/!!!!!)4:;)4#!!"#$%&"'"()*+,-(".)/'012&"'/3"'1"4#!9!<!!!
0.3
):4;)4#!!"#56=!"#$!*>27)%*!40*!"?@A!1,604*$!02'!2!',7)%*!+#$72%!'0#-B!.,11('*$/!&0,%*!
40*!"?@C!02'!2+!#5%,3(*!'0#-B!.,11('*$!'(-0!2'!&,%%!5*!.,'-(''*.!+*>4=!
0.2
!
!
0.1
D5%,3(*!?!E0#-B!.,11('*$'F!
0
!
0
1
2
3
4
5
6
G0*!%#''*'!-2+!5*!6$*24%H!$*.(-*.!5H!.*-*%*$24,+6!40*!1%#&!40$#(60!#+*!#$!7#$*!
M0
#5%,3(*!'0#-B'/!40*!.*1%*-4,#+!2+.!40*!)$*''($*!$,'*!#1!*2-0!5*,+6!'72%%!*+#(60!4#!5*!
,+!40*!$2+6*!&0*$*!40*!'426+24,#+!)$*''($*!$24,#!,'!-%#'*!4#!(+,4H=!!I4!,'!J*$H!
Figure 5.4: Stagnation
pressure ratio pt1 /pt0 (dashed line) and static pressure ratio p0 /p1 (solid line). Red,
= 1.66. Green,,7)#$42+4!4#!(+.*$'42+.!4024!2+!D5%,3(*!E0#-B!,'!,+!12-4!K('4!2!+#$72%!'0#-B!
= 1.4. Blue, = 1.3.
'42+.,+6!24!2+!2+6%*!4#!40*!1%#&=!!!
!

!
!
L%%!#1!40*!-02+6*!2-$#''!40*!'0#-B!42B*'!)%2-*!5*-2('*!40*!82-0!+(75*$!+#$72%!4#!
40*!'0#-B!,'!%2$6*$!402+!(+,4H=!!G0*!J*%#-,4H!+#$72%!4#!40*!'0#-B!.*-$*2'*'!&,40!
Figure 5.5: Oblique-shock diffuser
-#+'*3(*+4!,+-$*2'*'!,+!4*7)*$24($*!2+.!)$*''($*/!5(4!40*!J*%#-,4H!)2$2%%*%!4#!40*!
'0#-B!,'!(+-02+6*.F!

All of the change across the shock wave takes place because the Mach number corresponding to the velocity
component normal to the shock is larger than unity. The velocity normal to the shock decreases with
consequent increase in temperature and pressure; at the same time the velocity parallel to the shock is
unchanged.
M1n is given in terms of M0n by the same relation given for M1 as a function of M0 . The relations
obtained for the pressure ratios through a shock wave as a function of M0 are here function of M0n . The
advantage is that even for M0 very large, M0n can be made close to 1. Of course the condition for having
a weak wave is to have at least M0n = 1, or
sin()

1
M0

(5.8)

By choosing the wedge angle (or deflection angle) we can set the shock angle.
A series of weak oblique shocks, for each of which the Mn is near unity, hence all lying in the range of
small stagnation pressure loss, can yield an efficient diffuser.
3. Diffusers with internal contraction
One might ask why we do not just use a convergent-divergent nozzle in reverse as an inlet
39

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

!
!
!
!
!
"#$!%&!'%()$!%$!*)+,&!-.!"-$!/0!*1)!&2,)!+)32*%-$!'%()$!.-+!"#!2&!2!.4$5*%-$!-.!"-6!!
74*!"-$!52$!/)!,28)!53-&)!*-!#6!!91)!5-$8%*%-$!.-+!2!:)2;!-+!&-4$8!:2()!%&!<4&*!"-$!
=!#>!-+!
Aerospace! Propulsion
Lecture notes
!
!
?%$!!=!!"##$#%!@:)2;!5-,A+)&&%-$B!
70!51--&%$'!*1)!:)8')!2$'3)!@-+!8).3)5*%-$!2$'3)B!"#:)!52$!&)*!*1)!&1-5;!2$'3)6!
!
!C!&)+%)&!-.!:)2;!-/3%D4)!&1-5;&>!.-+!)251!-.!:1%51!*1)!"$!%&!$)2+!4$%*0>!1)$5)!233!
30%$'!%$!*1)!+2$')!-.!&,233!A*!3-&&>!52$!0%)38!2$!)..%5%)$*!8%..4&)+6!
!
!
!
!
!
!
!
!!
!
!
!
! E%..4&)+&!:%*1!%$*)+$23!5-$*+25*%-$!
! #$!%&!'%()$!%$!*)+,&!-.!"-$!/0!*1)!&2,)!+)32*%-$!'%()$!.-+!"#!2&!2!.4$5*%-$!-.!"-6!!
"
F$)!,%'1*!2&;!:10!:)!8-!$-*!<4&*!4&)!2!5-$()+')$*G8%()+')$*!$-HH3)!%$!+)()+&)!2&!
74*!"
Figure 5.6: Oblique shock wave
-$!52$!/)!,28)!53-&)!*-!#6!!91)!5-$8%*%-$!.-+!2!:)2;!-+!&-4$8!:2()!%&!<4&*!"
-$!
2$!%$3)*I!
=!#>!-+!
!
!
!
?%$!!=!!"##$#%!@:)2;!5-,A+)&&%-$B!
70!51--&%$'!*1)!:)8')!2$'3)!@-+!8).3)5*%-$!2$'3)B!"#:)!52$!&)*!*1)!&1-5;!2$'3)6!
!
!C!&)+%)&!-.!:)2;!-/3%D4)!&1-5;&>!.-+!)251!-.!:1%51!*1)!"$!%&!$)2+!4$%*0>!1)$5)!233!
30%$'!%$!*1)!+2$')!-.!&,233!A*!3-&&>!52$!0%)38!2$!)..%5%)$*!8%..4&)+6!
!
!
!
!
!
!
!
!
!
E%..4&)+&!:%*1!%$*)+$23!5-$*+25*%-$!
Figure 5.7: Is it possible to use a convergent-divergent nozzle in reverse as an inlet?
!
F$)!,%'1*!2&;!:10!:)!8-!$-*!<4&*!4&)!2!5-$()+')$*G8%()+')$*!$-HH3)!%$!+)()+&)!2&!
2$!%$3)*I!
This would work at one design Mach number, the one for which the isentropic area ratio between the
incoming supersonic flow and the sonic throat is exactly the as-built area ratio A1 /Athroat . Nevertheless,
during the acceleration to the design Mach number, it is not possible to establish fully supersonic flow in
the inlet without varying the geometry. To see this, imagine the inlet flying at M0 , lower than the design
Mach number. The flow will look as depicted in the right panel of Fig. 5.8. This is because at the lower
M0 the flow area that would decelerate isentropically to sonic at the throat is smaller than the built A1
The starting problem of internal-compression diffuser is schematically presented in Fig. 5.8. We consider
1
the development of the flow with increasing flight Mach number for a diffuser of fixed area ratio, A
At ,
between capture (diffuser inlet) and throat.
The ideal, shock-free operation (Fig. 5.8 (a)) would require a value of M0 such that7
A(M0 )
A1
=
.
At
A

(5.9)

7 A is the critical area of the diffuser. The expression of the area ratio of a diffuser is equivalent to the one for nozzles that
was studied in detail in previous courses and is equivalent to eq. 5.18. The variation of the area ratio with the Mach number is
presented in Fig.5.9

40

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 5.8: Schematics of internal-compression diffuser, showing (a) ideal isentropic diffusion from M0 through
unity to M < 1, (b) operation below the critical (starting) Mach number, (c) operation at the critical M0c , but
not started, and (d) operation at the critical M0c and started, with the shock positioned at the throat (from
Kerrebrock)
So for any smaller value of M0 ,

A1
A(M0 )
>
At
A
and considering that the flow is chocked at the throat, At = A , then
A1 > A(M0 )

(5.10)

(5.11)

This means that the streamtube that can be captured by the diffuser (A(M0 )) is smaller than the diffuser
inlet (A1 ); the rest of the flow into the frontal area A1 has to be somehow spilled. This spillage cannot
occur in supersonic regime since information is not able to travel upstream. Therefore a detached shock
forms ahead of the lip and the spill of excess mass flow occurs in the subsonic flow behind it. This situation
is shown in Fig. 5.8 (b). It has to be remarked that after the normal shock wave the stagnation pressure
decreases and the mass flow that has to be spilled is thus even higher. We have now to consider the
subsonic Mach number after the normal shock wave M0,1 When the flight Mach number, M0 , is increased
A(M0,1 )
1
, the normal shock will stand just at the lip, as in figure
to the critical value M0 c such that A
At =
A
5.8 (c). But in this position it is unstable and will move downstream if perturbed. Unfortunately, the
shock at the full flight Mach number is very lossy, and it is not practical to simply force the aircraft to
41

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

2.5
A1/A*

A/A*

A /A*

1.5

M=1

0.5

0
0

0.5

M0<MD

1.5
Mach number

MD

2.5

Figure 5.9: A/A vs. Mach number ( = 1.4).


continue accelerating to the critical condition (there may not even be enough thrust left to do it). For
a sufficiently high design Mach number there may even not exist a critical Mach number M0 c to ingest
the shock wave since the Mach number after the shock decreases with increasing M0 with an asymptotic
behaviour. What can be done is to manipulate the geometry to swallow the shock at a lower Mach number
and then reduce its strength.
This instability is easily understood by imagining that the shock is moved slightly downstream by some
disturbance. Since the Mach number decreases downstream with supersonic flow in the converging passage,
the perturbed shock will stand at a lower Mach number and will cause less stagnation pressure loss. The
choked throat will then pass more mass flow than what is captured by the lip, with a consequent net
outflow of mass from the convergent section of the diffuser. This requires a reduction in average density
that can be obtained only by discharging the shock downstream, thus starting the diffuser.
To achieve the best possible pressure recovery with the diffuser, once it is started, the back pressure would
be adjusted so that the shock stands at the throat, where the Mach number is smallest, as in figure 5.8
(d). This procedure should be performed carefully since the shock may accidentally pop all the way to
the front (an unstart, which is a violent transient event).
This is an adequate solution for modest values of the Mach number in the throat but for values of Mach
number in the throat higher than 1.5 the loss due to the shock at the throat is still excessive and a variable
geometry air intake is required.

5.4

Exhaust nozzles

As well as the inlet, the nozzle can range from very simple to quite complex. A simple fixed-area convergent form
usually suffices for subsonic aircraft except when jet noise suppression is required, while a complex variable-area
convergent-divergent device is essential for adequate performance in supersonic aircraft. In either case, the
functions of the nozzle are three:
Efficiently expand the high-pressure, high-temperature gases at the engine exhaust to atmospheric pressure, converting the available thermal energy to kinetic energy.
42

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Provide the required throat area able to swallow the mass flow coming from the turbine.
Try to recover pressure matching conditions at the exit of the engine (given that the pressure matching
condition is the optimum operation of the nozzle).
Throat Area
The required throat area is determined by conservation of mass. The mass flow rate at the compressor inlet
can be computed as shown in 3:
pt2 A2
(5.12)
m
= m2
RTt2
and if the throat (station 8) is choked,
pt8 A8
pt5 A8
m
=
=
RTt8
RTt5

(5.13)

Equating and using


pt5
pt2
Tt5
Tt2
we can calculate

A8
=
A2

(c t ) 1

(5.14)

(5.15)

m2

+1
2(1)

(5.16)

The ratios in the expression for A8 /A2 are determined by cycle analysis, as outlined previously, so from this
expression we can find the nozzle area ratio as a function of the engine parameters. As we have seen in our
discussion of the matching of components (3), once the nozzle area is set, the operating point of the engine
depends only on the turbine temperature ratio .
Exit Area The ratio of exit area to throat area required for ideal expansion can be found from the usual
compressible channel flow relations. Thus the exit Mach number is set by the stagnation pressure of the jet and
the ambient pressure (assuming pressure matching conditions)



pt9
pt9
1 2 1
= (0 c t ) 1
(5.17)
M9
=
= 1+
p0
p9
2
and the area ratio is then set by this Mach number,
+1
1 2 2(1)
M9
1+
A9
1
2

A8
M9
+1
2

(5.18)

Since pt9 /p0 involves 0 , M9 for this matched condition does depend on flight Mach number, and so does then
A9 /A8 . This is important for supersonic engines, as we discuss below.
These relations taken together serve to define the geometry of an ideally expanded nozzle. If the nozzle is
not ideally expanded, the behavior is more complex and will be discussed in the course on Rockets.
Effects of nozzle mismatching: Subsonic vs. supersonic
In this section, the effect of mismatching in nozzles is evaluated. We can use Eqs. (3.30) and (3.34) to
calculate the thrust of engines whose nozzles are respectively pressure-matched or truncated at the sonic point.
We illustrate this for two different engine designs, one subsonic and one supersonic:
43

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Case 1: M0 = 0.85 (0 = 1.1445)


Case 2: M0 = 2 (0 = 1.8)
For both cases, we take t = 6.25 and M2 = 0.6 (or
m2 = 0.8416). We also assume the compressor ratio is that
which gives maximum thrust in each case (c = t /0 ). The results obtained are shown in Table 5.1
M0
0.85
2

5.4609
3.4722

c
2.1844
1.3889

t
0.7831
0.8880

Athroat /A2
0.2658
0.7093

A9 /A2
0.2700
0.7190

(2 )matched
1.6635
0.9682

(2 )truncated
1.5108
0.7760

Table 5.1: Effects of nozzle mismatching


There is little effect (10%) on thrust when the nozzle is truncated at its throat for the subsonic nozzle, but
the effect is a 20% thrust reduction in the supersonic case. This fact is why such engines carry an adjustable
convergent-divergent nozzle (variable geometry).
Additional Requirements
The exhaust nozzle may have to meet a number of other requirements. Some are:
Variable area for afterburning, to increase throat area in proportion to the square root of the temperature
after afterburning

Figure 5.10: Schematics of two types of ejector nozzles in which secondary airflow is used to vary the expansion
ratio of the nozzle
Such nozzles have been built in a number of forms. At the top of Fig. 5.10 the relatively simple type used
on the F-111 and F-15 engines is displayed. These nozzles have top flight Mach numbers of the order of
2. The bottom of the figure shows the nozzle type used on the SR-71, which reaches 3.5 or thereabouts.
Altitude compensating nozzles

This kind of nozzles are able to maintain their efficiency over a wide range of altitudes. An example of
this is shown in Fig. 5.11.
In these nozzles, commonly known as aerospike, instead of firing the exhaust gases through a small hole
in the middle of a bell, an aerospike engine avoids this random distribution by firing along the outside
edge of a wedge-shaped protrusion, the spike. The spike forms one side of a virtual bell, with the other
side being formed by the outside air, thus the name aerospike. The idea behind the aerospike design
is that at low altitude the ambient pressure compresses the wake against the nozzle. The recirculation in
the base zone (central region) of the wedge can then raise the pressure there to near ambient. Since the
pressure on top of the engine is ambient, this means that base gives no overall thrust due to pressure (but

44

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 5.11: Aerospike or plug-nozzle (Altitude compensating nozzles)

Figure 5.12: Aerospike nozzle plume at different conditions


it also means that this part of the nozzle doesnt lose thrust by forming a partial vacuum, thus the base
part of the nozzle can be ignored at low altitude). This operation is depicted in Fig.5.12.
As the spacecraft climbs to higher altitudes, the air pressure holding the exhaust against the spike decreases, but the pressure on top of the engine decreases at the same time, so this is not detrimental.
Further, although the base pressure drops, the recirculation zone keeps the pressure on the base up to a
fraction of 1 bar, a pressure that is not balanced by the near vacuum on top of the engine; this difference
in pressure gives extra thrust at altitude, contributing to the altitude compensating effect.
Noise suppression

As we shall see in 12.4 on noise, the principal way to decrease jet noise is to lower the jet velocity for
a given thrust. The turbofan is the most efficient way to do this, but for some applications it is not
practical to use a fan. In this case there is the desire to increase the mass flow rate of the jet, by mixing
in additional air that lowers the velocity for a given total momentum. One way to do this is the lobed
mixer (Fig. 5.13)
Lobed mixers are also used sometimes in turbofans to help equalize the core and bypass stream velocities,
and hence increase performance (by 3-4%). The improvement is, however, less than if the velocity changes
were isentropic.

Thrust-vectoring, as for advanced VTOL aircraft.

45

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 5.13: Schematics of lobe mixers

Figure 5.14: F-18 High Angle of Atack Research Vehicle

46

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Principles of Compressors and Fans

Compressors are fluid dynamic devices, i.e., they depend on fluid accelerations to compress and expand gases, in
contrast to positive displacement devices such as the familiar piston engine. The moving blades of a compressor
exert a force on the fluid by virtue of their motion, resulting in an added velocity of the fluid, and consequently
an increase in its energy in stationary coordinates. The force is dependent on the blades velocity, roughly as
the square, that makes this a dynamic machine. In contrast, the energy per unit mass added to the fluid in the
positive-displacement piston-cylinder mechanism is nearly independent of the speed of the piston, depending
only on the volumetric compression ratio.

Positive displacement

Fluid dynamic

Figure 6.1: Positive displacement devices vs. fluid dynamic devices.


Large aircraft engine compressors or fans are mainly of the axial-flow type, with rotating rows of rotor blades
intercalated with stator vanes. A cross-sectional view through the axis is shown in Fig. 6.2.

Figure 6.2: Cross-sectional view through the axis of an axial compressor. IGV denotes the inlet guide vanes; R
denotes the rotor blades; S denotes the stator vanes.
Both rows of blades function to deflect the flow in the tangential direction, the rotors adding angular
momentum in the direction of the rotation, the stators removing it. Because of their motion, the rotating
blades do work on the flow, increasing its energy. The stator vanes in contrast only diffuse the flow, exchanging
momentum for pressure rise.
If we make a cut as indicated in Fig. 6.2, we would obtain a surface, whose cross section would be formed
by a circumferential array of blades. To simplify the rotational symmetry of the actual blade rows, we think
47

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

of a planar cascade, extending to infinity in both directions (Fig. 6.3). In this limit, we leave out the radial
motion of the fluid, and the rotor blades then have a constant velocity (while the stators are stationary).

Figure 6.3: Cross section of the blades.


To describe the effect of the blades, velocity diagrams are used (see Fig. 6.4), which show how the velocities
change across the rotor and stator rows, and in particular the effects of the relative motion of the two. The
subindex a stands for the station before the rotor blades, b stands for the station after the rotor and before
the stator and c stands for the station after the stator. The apex refers to velocities and angles in the
rotating reference frame which are also represented with dotted lines.

Figure 6.4: Velocity diagrams showing how the velocities change across the rotor blade and stator vane rows.
If the velocities at c are the same as at a, we call the stage (rotor + stator) a repeating stage.

6.1

Eulers equation

The energy exchange takes place in the rotor blade rows. The azimuthal force on a blade F is equal to the rate
of change of azimuthal momentum per blade:
F =m
per blade (vb va )
48

Bioengineering and Aerospace Engineering Dept.

(6.1)
v. 2015-2016

Aerospace Propulsion

Lecture notes

where v is the azimuthal (or tangential) component of V at any point.


The power delivered to the fluid by a blade is the force times the blade velocity and this must equal the
change in fluid energy per blade, across the rotor, so
Power
Power per blade

= F r
=

rm
perblade (vb va )

= m
perblade Cp (Ttb Tta )

(6.2)
(6.3)
(6.4)

leading to
Cp (Ttb Tta ) = r(vb va )

(6.5)

This is the famous (and important) Euler Equation, for a 2D cascade of rotor blades.
The energy exchange across the rotor is most generally understood as follows. We focus on a streamtube
passing through the rotor, with a mass flow m,
as shown in Fig. 6.5.

Figure 6.5: Sketch of a streamtube passing through the rotor.


The torque on the rotor, T , due to this stream tube is
T = m(r
b vb ra va )

(6.6)

Power = T = m(r

b vb ra va )

(6.7)

Power = mC
p (Ttb Tta ),

(6.8)

Cp (Ttb Tta ) = (rb vb ra va )

(6.9)

Cp (Ttb Tta ) = (rb )2 .

(6.10)

and the power is


Equating this to the total enthalpy rise of the fluid

This is the Euler Equation in its general form. The left hand side is replaced by the more general form
htb hta (the total enthalpy rise in the streamtube across the rotor) when the gas cannot be modeled as being
ideal.
The Euler Equation applies to other geometries than the axial flow one, e.g. the centrifugal or radial
compressor
In this case va = 0 and if the vanes are radial, vb r so we have
For the axial compressor, rb ra , and we have
Cp (Ttb Tta ) r(vb va ),
49

Bioengineering and Aerospace Engineering Dept.

(6.11)
v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 6.6: Sketch of a centrifugal compressor


and then

Ttb
(r)2
1
Tta
Cp Tta

vb va
r

R Ta (r)2
=
Cp Tta RTa

vb va
r

vb v a
r


,

(6.12)

leading to
Ttb
Ta 2
1 ( 1)
M
Tta
Tta T

(6.13)

where MT is the tangential Mach number of the blade motion.

6.2

Velocity triangles

In the previous section we discussed the basic mechanisms of energy exchange in compressors and drew some
simple velocity triangles to show how we go from the stationary coordinate system to one in the moving blades.
In more detail, the velocity triangle for the flow before the rotor and after the rotor is shown in Fig. 6.7 (note
that the axial velocities wa and wb are not necessarily the same as in the sketch).
Now we can write
va
vb

=
=

wa tan a
wb tan b = r

(6.14)
wb tan b0

(6.15)

where b0 is the flow angle in the rotating coordinate system of the rotor blades. The advantage of the latter
form for vb is that it is b0 which is nearly set by the shape of the blades, i.e. b0 is nearly constant. The difference
between the flow angle and the angle of the chord line at the trailing edge of the blades is called the deviation,
and is usually a small number of degrees.
Taking advantage of this small deviation, we can write the Euler equation as


( 1)MT2
wb tan b0
wa tan a
Ttb
=1+
1

,
(6.16)
Tta
r
r
1 2
1+
Ma
2
50

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 6.7: Velocity triangle.


where Ma is the flow Mach number upstream and MT is the tangential Mach number of the blade motion.
Since the two flow angles expressed this way are nearly constant, the temperature ratio becomes a function
primarily of two variables, the tangential Mach number MT and the ratio of the axial flow velocity to the blade
speed. For the usual case of b0 > 0, a > 0 we get a functional dependence as shown in Fig. 6.8.

Figure 6.8: Temperature ratio vs. ratio of the axial flow velocity to blade speed for various tangential Mach
numbers.

6.3

Isentropic efficiency and compressor map

In reality, compressors provide a lower stagnation pressure raise than the ideal one for a given work input,
/(1)
c = c
, due to entropy increase by viscosity and other factors. Conversely, this means that we need to
provide more work than the ideal (i.e., we need to increase more the total temperature) to achieve the same
pressure ratio. If we consider a given stagnation pressure raise, the ratio of the ideal temperature increase to

51

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

the actual, necessary one is called the (isentropic) compressor efficiency,

(Tt )ideal
c =
=
(Tt )actual

Tta

ptb
pta

! 1

Ttb Tta

Tta

ptb
pta

! 1

Ttb
1
Tta

(6.17)

Since c < 1, we have c = [1 + c (c 1)] 1 < c1 .


We usually plot Eulers equation in terms of the pressure ratio rather than the temperature ratio, and include
superimposed lines of efficiency, so that the map looks like in Fig. 6.9.

Figure 6.9: Performance map for a typical high-pressure-ratio compressor (refer to the caption of Fig. 3.2.

52

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

7
7.1

Lecture notes

Compressor Blading, design and multi-staging


Diffusion factor. Stall and surge

By decreasing a and b0 , or even making them negative, we can increase Ttb /Tta . What limits this increase?
The answer is best given in terms of a Diffusion Factor, which describes the tendencies for the boundary layer
to separate under the influence of the pressure rise in the blade passage, as sketched in Fig. 7.1.

Figure 7.1: Schematic diagram of velocity distributions on suction and pressure surfaces of a blade, showing the
diffusion from a maximum velocity Vmax to the final velocity Vb and the resultant thickening of the boundary
layer on the suction surface.
The critical region is the suction surface (extrados) of the blade. This region feels a pressure rise due to
the decrease of V 0 from Va0 to Vb0 , (the prime is a reminder that these velocities in the relative frame) and also
due to the acceleration, followed by deceleration on the suction side. For compressors and fans, we define the
Diffusion Factor as
V0
|vb va |
D = 1 b0 +
,
(7.1)
Va
2Va0
which is a crude but effective way to account for these two flow deceleration components. Here, = c/s is the
solidity of the rotor, where c is the chord of the blades and s is their spacing (in the peripheral direction). A
value of about 0.5 for D is the upper limit for good efficiency.
As the angles of atack increase in the compressor a point is reached when the flow begins to separate from
the blades, with serious consequences. Large incidence angles occur when the ratio (axial velocity)/(blade
speed) is low, i.e., for a fixed rotational velocity, at reduced flow, and hence at increased pressure ratio. All
compressors have a fairly well defined stall line that runs diagonally from low flow and low c , to high values
of both. Operation must be restricted to the region below and to the right of this line. The operational line
for the compressor runs roughly parallel to this stall line. As we saw in 3, the reason for the compressor to
be restricted to its operation line has to do with flow continuity between it and the turbine and nozzle. It is a
fortunate fact that the line thus defined tends to correspond to a constant blade incidence, which, when chosen
properly, is close to that which optimizes blade performance, and so the operating line is more or less the peak
compressor efficiency line.
53

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

The phenomena that occur at and beyond stall are complex, dynamical and highly nonlinear. A detailed
understanding of these effects has only emerged in the last decades, and remedial measures based on this
understanding are still under evaluation.
In most axial compressors, incipient separation first leads to the Rotating Stall phenomenon: sections of
the stalling rotor then operate in deep stall, with almost zero flow, while the rest carries normal or high flow
per unit frontal area. These regions move backwards in the rotating frame at 0.4 0.6r, so that, when
observed from rest they rotate forward, but only at a fraction of the rotor speed. The reason for the bimodal
flow distribution is that adjacent streamtubes become unstable with respect to flow exchange: if the rotor as a
whole is near stall conditions, and one streamtube loses some flow and diverts it to its neighbors, the streamtube
with less flow goes into stall and loses even more flow, while the neighbors gain flow and remain stable. The
reason the stall cells move backwards relative to the rotor is an elaboration of the same argument: if one flow
passage stalls, the swirling incoming flow is re-routed such that the passage ahead of the one stalled sees a more
axial flow, while the one behind sees a larger incidence angle. The stall moves to this trailing passage, and the
stalled passage clears.
Rotating stall, since it moves rapidly about, tends to average out and, aside from high-frequency excitation,
may not be dynamically significant. On the other hand, the net compressor performance drops strongly, and
in addition, it is not easy to reverse once started, except by stopping and re-starting the engine. Detecting
rotating stall and instituting the appropriate control reaction is very important. If the engine controls simply
detect a loss of pumping performance (pressure loss) they may react by increasing fuel flow, which combined
with the reduced airflow, may lead to overheating and burnout.
Under some conditions having to do mainly with the ratio of flow inertia to flow passage capacitance, the
complete pumping system (compressor plus choked turbine nozzles) can enter a global oscillation, called
Surge. Unlike rotating stall, surge involves deep oscillations or even reversals of the whole flow through the
compressor, and can be mechanically destructive (certainly quite noticeable, in the form of loud, repeated bangs,
accompanied by flame ejection from both ends of the engine).
When stall is reached, the engine may go into either Rotating Stall or Surge. The detailed mechanisms that
determine which of the phenomena will prevail are encapsulated in Greitzers B parameter
r
r Vp
B=
(7.2)
2a Vc
where a is an average speed of sound in the burner, Vp is the plenum volume (mainly the burner volume) and
Vc is the volume in the compressor flow passages. In a single-stage compressor, values of B below 0.8 lead to
rotating stall,
while higher values lead to surge. For multistage (N ) compressors, Bcrit is lower by somewhere

between N and N . One favorable aspect of surge (as opposed to rotating stall) is that it can usually be
cleared by simply reducing fuel flow (or, in a test stand, opening the downstream throttle).

7.2

Compressor blading and radial variations.

In all of the discussion so far the blade speed has been taken as a parameter. In fact the blade linear speed
varies with the radial location in the compressor, so the designs that we have discussed are applicable only at
one radius in any real compressor. In practice it is usual to begin a compressor design with such a treatment,
called a mean line design carried out at some mean radius. But the effects of the blade speed variation with
radius are important, so we must be aware of the constraints they impose on the design.
Normally it is desirable for the total pressure (or temperature) ratio of a rotor blade row to be approximately
constant over the radial length of the blade, so that the outlet airflow has uniform pressure. From the Euler
equation,
(rb vb ra va )
Ttb
1=
.
(7.3)
Tta
Cp Tta

54

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Subjects: Compressor Blading; Design; M ulti-Staging


Compressor blading; Radial variations
In all of the discussion so far the blade speed has been taken as a parameter. In fact the
Lecture notes
blade speed varies with the radial location in the comp!"##$!%&#$&'("&)*"#+,-#.&'(/'&0"&
have discussed are applicable only at one radius in any real compressor. In practice it is
1#1/2&'$&3",+-&/&4$56!"##$!&*"#+,-&0+'(&#14(&/&'!"/'5"-'%&4/22"*&/&)5"/-&2+-"&*"#+,-.&
carried out at some mean radius. But the effects of the blade speed variation with radius
are important, so we must be aware of the constraints they impose on the design.

Aerospace Propulsion

Figure 7.2: Cross-section of an axial compressor showing the variation of the radius with axial distance.
As a simple example, suppose that va = 0 (no inlet guide vanes). We see the condition that the total temperature
Normally it is desirable for the pressure (or temperature) ratio of a rotor blade row to be
ratio remains constant with radius requires

approximately constant over the radial length of the blade,


so that the outlet airflow has
const
.
(7.4)
b vb = const. ; vb =
uniform pressure. From the Eulerrequation
rb
r1 1 )Vortex flow (Fig. 7.3). Now let us draw the
Tt 2should generate
( r2 2 a Free
This implies that the rotor blade row
1
=
velocity triangles for tip and hub (seeTFig. 7.4), further
cpTt1assuming rH /rT = 1/2, vb (rH ) = rH (that is that
t1
b0 = 0), and
rT
Supposing for simplicity that 1 = 0wa(no
guide vane),
= winlet
. we see this condition requires(7.5)
b = const. =
2
const
Hence, for this example,
r2 2 = const.; r 2 = r
H
T
vb (rT ) =
= r2 .
(7.6)
2
4
This implies that the rotor blade row should generate a Free Vortex in the flow

Now let us draw the velocity triangles


and flow
hub,(vassuming
rH/rT = !, and w1 = w2 =
Figure 7.3: for
Freetip
vortex
1/r)
const = rT /2.

The approximate blade shapes as sketched are determined by the condition that the leading and trailing
edges are aligned with the flow. We see that the blades are strongly twisted from hub to tip.
Now let us see what these variations imply for the Diffusion Factor, D (Remember that V 0 are velocities
relative to the rotor blade).
V0
|vb va |
D = 1 b0 +
(7.7)
Va
2Va0
55

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 7.4: Sketch of the hub and the tip profiles and corresponding velocity triangles.
From the geometry, for the TIP and taking va = 0 as in the previous example,
|vb va | =
Va0
Vb0
DT

rT /4
r

(7.8)

rT

(7.9)

1
= 1.12rT
4
s 
 2
2
1
3
= rT
+
= 0.901rT
2
4
0.116
= 0.195 +
T
=

1+

(7.10)
(7.11)

and for the HUB


|vb va | =
Va0
Vb0

DH

rH = rT /2
p
p
rH (2) = rT (2)/2
rH = rT /2
0.353
0.293 +
H

(7.12)
(7.13)
(7.14)
(7.15)

The solidities at hub and tip, H = cH /sH and T = cT /sT are of course design choices, but since the
spacings sH and sT are related by
sH
rH
=
(7.16)
sT
rT
H
cH sT
cH rT
=
=
(7.17)
T
cT sH
cT rH
Let us choose T = 1, CH /CT = 1, then H = 2, leading to
DT

DH

0.195 + 0.116 = 0.311


0.353
0.293 +
= 0.470
2

(7.18)
(7.19)

Both of these are acceptable from the viewpoint of losses, but DH barely so. It is generally true that the
hub section of the blade is limiting from the viewpoint of diffusion.

7.3

Multi-staging and flow area variation

The reason behind multi-stage compressors is the limited pressure ratio that can be achieved with a single stage.
Typically, several rotor-stator stages are stacked up to produce the required compression.
56

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Because the density increases, the axial velocity will decrease unless we reduce the flow area accordingly.
Axial deceleration in a compressor is severely limited, as it can stall the airfoils of the blades. Thus, the flow
area must decrease as we go through the compressor to maintain near-constant axial velocity. We can choose
to effect this area variation by decreasing the tip radius, by increasing the hub radius, or something in between,
like keeping a constant mean radius. All these solutions have both advantages and disadvantages.
The constant outer diameter design maintains the blade velocity at the maximum for the entire length of
the compressor, hence gives the highest pressure ratio for a given number of stages. But because the length of
the blades becomes small at the final stages of the compressor, leakage through the tip clearance is most serious
for this design choice.
The constant inner diameter design minimizes the tip clearance problem and also yields lower stresses in the
discs of the last stages, but requires more stages for a given pressure ratio (since the lower-radius blades have
a lower speed and exert less work than higher-radius ones).

7.4

Mach Number Effects

The performance of compressor blades deteriorates once the relative inlet Mach number exceeds 0.70.8, since
on the extrados of the airfoil the accelerated flow typically reaches supersonic conditions, creating shock waves
and thicker boundary layers which decrease the total pressure of the flow, incurring in efficiency losses. Since
the linear velocity is maximal at the blade tip, it is also there where the Mach number is higher.
There are however reasons that encourage us to go beyond this limit: a higher Mach number implies a higher
mass flow per unit area, so that the compressor is more compact. Second, a higher Mach number is caused
by a higher blade velocity, which enables more work to be deposited into the flow, and hence higher pressure
ratios at the cost of lower compressor efficiency (there is of course a trade-off here, as stronger shock waves
will ultimately cancel the increased pressure gain). Note that, while on-ground compressors aim to maximize
efficiency, aircraft compressors also have to consider the work exerted by each stage to reduce the number of
stages and result in a lighter device (even at the cost of efficiency, which is usually reduced a bit).
Hence, in many circumstances, compressors are designed to operate in a transonic regime in their first stages
(later stages have lower Mach numbers since temperature and sound speed increase downstream). Typical
values of first-stage blade tip relative Mach numbers can reach as high as 1.5 in modern engines. As an example,
for an axial Mach number 0.7 and blade tangential Mach number MT 1.3, we have at the tip
q
0
(7.20)
Ma,T
= Ma2 + MT2 = 1.48
so the tips of the blades see supersonic flow. For rH /rT = 0.5, at the hub

0
Ma,H
= 0.49 + 0.42 = 0.95

(7.21)

so the hub is at a very high subsonic speed. Just as for inlets, hub and tip therefore require quite different
diffusion techniques. To alleviate the effects of the high relative Mach number, very thin blades are used to
reduce their blockage, with relative thicknesses of only a few percent (see Fig. 7.5). At the hub, instead, a
blading that is the equivalent of a subsonic inlet is required.
The shockwaves that form lead to a rise in entropy of the flow, but it does not necessarily lead to excessive
losses if the blades are designed carefully.

7.5

The Polytropic Efficiency

What is almost nearly constant among stages of a multi-stage compressor is the limiting isentropic efficiency
for a small, differential-like compression, which is called the Polytropic efficiency:

poly
57

dpt
(dht )s
RTt d ln pt
1 d ln pt
t
=
=
=
=
dht
Cp dTt
Cp dTt
d ln Tt

Bioengineering and Aerospace Engineering Dept.

(7.22)
v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 7.5: Sketch of the supersonic flow on the first rotor row and the system of shockwaves and expansion
waves.
(remember that dh = T ds + (1/)dp). If we assume poly remains constant along the compression path, this
definition can be integrated to

 1
1
Tt,out
pt,out poly

=
c = c poly .
(7.23)
Tt,in
pt,in
Inserting this into the definition of the global Isentropic Efficiency allows one to compute it in terms of poly
and the finite pressure or temperature ratio:
1

c 1
c 1
c
=
1
c 1

c poly 1

(7.24)

As an example, take a multi-stage compressor with an overall pressure ratio c = 16 and assume the
polytropic efficiency is constant and equal to 0.9 in all stages. Its isentropic efficiency (ideal work required
divided by actual work) is then calculated to be c = 0.856. This is less than poly , the small-increment efficiency,
because the last stages of the compressor receive gas that is hotter than it should, due to the inefficiencies in
the previous stages, and hence more work is needed to compress it. It could be mentioned here that a similar
argument can be made for turbines, and in that case one finds poly < t , because the extra thermal energy due
to inefficiencies of the early stages is now available for conversion to work by the latter stages.
Likewise: the global isentropic efficiency of a compressor with N stages, each with an isentropic stage
efficiency s (close to poly ) is not sN (actually, sN > c ).

58

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion
Starting and Low-Speed Operation

7.6

Lecture notes

Because
the Low-Speed
density variationOperation
through the compressor is much less at low speed
Starting
and

conditions, the compressor develops adverse blade loading situations in both the inlet and
Since the density variation through the compressor is much less at low speed conditions, but it is designed for
the outlet stages at low operating speeds. As shown in the sketch, the axial velocity in the
near-cruise conditions, the compressor develops adverse blade loading situations in both the inlet and the outlet
inlet stages tends to be lower, and that in the outlet stages higher, relative to the blade
stages at low operating speeds. As shown in Fig. 7.6, the axial velocity in the inlet stages tends to be lower,
speed, than at design, so the front stages tend to be stalled and the rear ones to
and that in the outlet stages higher, relative to the blade speed, than at design, so the front stages tend to
!"#$%&#''()*+,#-*&./0-*#1*%#22#34'1*15*%0-#6$*.*35&780--58*1,.1*"#''*5708.10*-.1#-2.3158#'9*
be stalled and the rear ones to windmill. This makes it difficult to design a compressor that will operate
over a wide range of speeds.
satisfactorily over a wide range of speeds.

Figure 7.6: Sketch of the velocities at design and low-speed operation.


The solution to this problem has taken two forms: one is to divide the compressor into
The solution
to this problem has taken two forms: one is to divide the compressor into two or even three
1"5*58*0:0$*1,800*!-755'-(;*5708.1#$6*.1*%#22080$1*-700%-;*1,0*51,08*#-*15*4-0*-1.158*<'.%0-*
spools, operating
at different
theflow
other
is to use
stator
blades
of modern
variablehigh
angle,
to adjust the flow
of variable
angle, to speeds,
adjust the
direction
at low
speeds.
Most
pressure
direction at low
speeds.
modern high pressure ratio engines use both. Also, intermediate bleeding of part
ratio
enginesMost
use both.
of the air flow is used to control this and other phenomena.
It is important to realize that the spools are all free-spinning, and seek their own equilibrium rotational
speeds in response to the forces on the blades and the torque exerted on the associated turbine. It is an
interesting exercise to show that the equilibrium of the spool rotation speed is, in fact, stable, i.e., the result of
an increased spin rate (at fixed axial flow rate) is a general increase in blade angles of attack, hence an increase
in torque that tends to restore the original spin rate.

59

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Turbines. Stage characteristics. Degree of reaction

Turbines behave according to the same principles of dynamic energy exchange as compressors. However they
differ from compressors in some important ways. First, since they extract energy from the flow rather than
adding it, the pressure drops throughout a turbine. This leads to different fluid mechanical limitations than for
compressors. Second, because of the thermodynamic requirements, they operate in gases at high temperatures,
so that materials and cooling requirements are of central importance. We will begin by exploring the fluidmechanical energy exchange in the turbine.

8.1

Eulers Equation

The exchange of energy is described by the Euler equation, just as for compressors. We will discuss here only
axial flow turbines such as are generally used in aircraft engines, but it is important to realize that turbines
come in a wide variety of types, from extremely large hydraulic turbines for hydroelectric power generation to
the familiar lawn sprinkler.
Axial flow turbines use a stator blade row (usually called a nozzle row) followed by a rotating blade row
that extracts the energy (often called the buckets). The nozzles expand the flow to a high velocity and turn
it, imparting an angular momentum to the flow. The buckets generally further expand the flow, and turn it
back toward the axial direction. The combination of the two blade rows is called a Turbine Stage. We use
the alphabetical notation or the different stages to avoid confusion with the numerical station notation for the
engine. Station a is equivalent to station 4 (turbine inlet) in the engine layout, and for a single stage turbine,
station c should be equivalent to station 5 in the engine. A general stage layout is indicated at the top and two
specific blade types at the bottom of Fig. 8.1. As it can be seen in Fig. 8.1, the spacing between two stator
blades resembles the shape of a nozzle. Due to their geometry, this nozzles are operating at chocked conditions
when the engine is working at full power.
From the Euler equation, for the general situation,
Cp (Ttc Ttb ) =
Ttc
=
1
Ttb

(vc rc rb vb )
rc
rb
(vb vc )
Cp Ttb
rc

(8.1)
(8.2)

Let us assume rb /rc = 1, and also that vc = 0. The latter is desirable if the rotor is the last stage of a
turbine, in order to avoid swirl in the flow entering the jet nozzle.
1

Ttc
rb
= 1 t =
vb
Ttb
Cp Ttb

(8.3)

numbers and the flow angles indicated in the diagram, with vb = Vb sin b =
Writing this in terms of Mach

RTb Mb sin b , and rb = RTb MT ,


1 t =

( 1)MT Mb sin b
1 2
1+
Mb
2

(No exit swirl)

(8.4)

Here MT is the tip Mach number of the blades and Mb is the flow Mach number. From this we can see that we
want large MT and large Mb sin b for large work per stage.

8.2

Degree of Reaction

As for the compressor there are many possible design choices for the turbine. One key choice is the relative
amounts of pressure drop in the nozzles and buckets. This is best characterized in terms of the Degree of
60

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 8.1: Construction of velocity triangles for a turbine stage, with typical composite diagrams for a 50%
reaction stage and an impulse stage.
Reaction, which is defined as the ratio of static enthalpy change in the rotor to that in the rotor plus stator
(stage). For the turbine, the degree of reaction is:
R=

hb hc
(hb hc ) + (ha hb )

(8.5)

We now use energy conservation in the nozzles station:


1
ha + Va2
2

1
hb + Vb2
2

(8.6)

1
hc + Vc2 rvc
2

(8.7)

and the rothalpy conservation in the moving frame:


1
hb + Vb2 rvb
2

This equation can also be obtained from the consevation of total enthalphy of the flow in the relative frame,
but in that case you must be sure to include all the potential energy terms at both sides of the equation (in
particular the potential energy due to the centrifugal forces8 ).
1
1
hb + Vb02 (r)2
2
2
8 In

61

1
1
= hc + Vc02 (r)2
2
2

(8.8)

our particular case, we are assuming rb = rc and therefore this terms disappear.

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Vc!

Vb

w = Vc = V a

Vb!
r
vb

Figure 8.2: Velocity triangle with constant axial velocity w = Va = Vc .


Replacing eq. (8.6) and eq. (8.8) in the definition of the degree of reaction (eq. (8.5)), we can obtain:
R=

Vc02

Vc02 Vb02
Vb02 + Vb2 Va2

(8.9)

For a general degree of reaction, but still with axial stage intake and exhaust, and with constant axial
velocity (Va = Vc = w), the velocity triangle is shown in Fig. 8.2, and we have
Vb2 Va2

Vc02

Vc02
Vb02
Vb02

vb2 ,

w + (r) ,

(8.11)

w2 + (vb r)2 ,

(8.12)

2rvb

(8.10)
2

vb2 .

(8.13)

Returning to eq. (8.9)


1R=
so that

vb2
vb
Vb2 Va2
=
=
02
2
02
2
Vc Vb + Vb Va
2rvb
2r

vb
2r
and using this to eliminate Mb sin b in the expression for the temperature ratio
R=1

1 t =

( 1)MT2
2(1 R).
1 2
1+
Mb
2

(8.14)

(8.15)

(8.16)

The value of MT is limited by the stresses and the temperature of the blade. For a given value of Mb , the
temperature drop in the turbine (hence its work per unit of mass flow, is related to R. As R increases, the work
decreases.

62

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

An alternative representation commonly usded is in terms of the so-called flow and power coefficients, and
. The definitions are

w
r
ht
vb vc
=
(r)2
r

(8.17)
(8.18)

For zero exit swirl, vc = 0, we can also put


=

Mb sin b
vb
=
,
r
MT

(8.19)

so that

or = 2(1 R)
(8.20)
2
which very directly shows how the stage power decreases with the degree of reaction. In terms of flow coefficient
and stator exit angle,
1
(8.21)
R = 1 tan b
2
!"#$#%&'#(()&)#*+$%&,*%,--%.#%/#,0)-1%2)$3,-)4#0%(/'5%,%$&,-#0%2#-'&)+1%+/),*6-#%)*%
These
coefficients can all be readily visualized from a scaled velocity triangle in which the blade speed r is
taken
to be unity (Fig. 8.3).
7")&"%+"#%.-,0#%$8##0%!/%)$%+,9#*%+'%.#%3*)+1:%
R=1

%
%
%
%
%
%
%
%
;.%
%
".%
%
%
%%%%%%%%%%%%%%%%#% %
%
%
!".%
%
%
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<%
%
%
%
%
%
%
%%%%%%%%%$=>%%%%%%%%?%
%
%
%
%
%
%
%
%
%
%
%
%
%
%
%%%%%%%%$%
%
@/,7)*6%+"#%2#-'&)+1%+/),*6-#$%('/%?A%<=>%B&,--#0%CDE%/#,&+)'*F%,*0%('/%?AD%B&,--#0%
Figure 8.3: Scaled velocity triangle
G583-$#FH%%
The velocity triangles for R = 1/2 (called 50% reaction) and for R = 0 (called Impulse) are sketched in Fig.
8.4.
In the 50% reaction turbine the enthalpy drop in the moving blades equals that in the nozzles. In the
impulse turbine there is no enthalpy (and thus pressure) drop in the buckets. It all takes place in the nozzle
guide vanes.
In designing the turbine we have some latitude in choosing the degree of reaction R. The tangential velocity
of the blades, hence MT , is limited by the strength of materials at high temperature, so it may seem that we
would like to use small R to maximize the temperature drop per stage. Nevertheless, the efficiency of the turbine
decreases as R is decreased from R = 1/2 toward R = 0. The boundary layers, in fact, have more tendency to
separate in the moving blades for R = 0 than for R = 1/2 because for R = 0 there is no pressure drop as the
%
flow is deflected.

G*%+"#%CDE%/#,&+)'*%+3/.)*#%+"#%8/#$$3/#%0/'8%)*%+"#%5'2)*6%.-,0#$%#I3,-$%+",+%)*%
+"#%*'44-#$H%7")-#%)*%+"#%)583-$#%+3/.)*#%+"#/#%)$%*'%8/#$$3/#%0/'8%)*%+"#%.3&9#+$J%
63
Bioengineering and Aerospace Engineering Dept.
v. 2015-2016
G+%,--%+,9#$%8-,&#%)*%+"#%*'44-#$J%
%
G*%0#$)6*)*6%+"#%+3/.)*#%7#%",2#%$'5#%-,+)+30#%)*%&"''$)*6%+"#%0#6/##%'(%
/#,&+)'*%?J%%%!"#%+,*6#*+),-%2#-'&)+1%'(%+"#%.-,0#$H%"#*&#%K!H%)$%-)5)+#0%.1%+"#%
$+/#*6+"%'(%5,+#/),-$%,+%")6"%+#58#/,+3/#H%$'%)+%5,1%$##5%+",+%7#%7'3-0%-)9#%+'%

Aerospace Propulsion

Lecture notes

Figure 8.4: Velocity triangles for R = 1/2 (50% reaction) and R = 0 (impulse).

8.3

Radial variations

Just as for the compressor, the requirement for a constant temperature drop from hub to tip across the turbine
flow path, places constraints on the degree of reaction. Since MT is proportional to radius, it is larger at the
tip than at the hub of the blades, so if Mb is about constant, R must decrease from the tip to the hub. Thus
typically if the degree of reaction is near 1/2 at the tip, it may be considerably smaller at the hub.

8.4

Rotating blade temperature

Another matter of interest is the temperature of the rotating blades, since the hotter they are the less stress
they can tolerate. As we shall see, the stagnation temperature relative to the moving blades is not equal to that
entering the nozzles.
0
Let Ttb
be the stagnation temperature relative to the rotating blade
0
Ttb
= Tb +

from which

0
Ttb
=1+
Ttb

V 02 Vb2
Vb02
= Ttb + b
2Cp
2Cp

(8.22)

Vb02 Vb2

(8.23)

2Cp Tb

!
1 2
Mb .
1+
2

Now, from velocity triangles,


Vb2
Vb02
After some algebra we obtain:

w2 (1 + tan2 b )

(8.24)
2

(8.25)

Vb02 Vb2 = r(r 2w tan b )

(8.26)

w + (w tan b r)

We now recall that the degree of reaction can be written as


R=1

1 w
tan b ,
2 r

(8.27)

so we can eliminate the term w tan b . The result is


Vb02 Vb2 = (r)2 (3 4R),
and substituting into eq. (8.23),

64

0
1 2
Ttb
=1
MT
Ttb
2

3 4R
.
1 2
1+
Mb
2

Bioengineering and Aerospace Engineering Dept.

(8.28)

(8.29)

v. 2015-2016

Aerospace Propulsion

Lecture notes

0
So we find that Ttb
is generally lower than Ttb = Tta , reflecting the fact that the blade is running away
from the stator flow, but also that it increases as R increases. Therefore for a given blade speed the impulse
turbine will have a lower blade temperature than the 50% reaction turbine. This just introduces one more factor
in the turbine design. To see the magnitude of this effect, take MT = Mb = 1 and = 1.4. Then, from eq.
(8.29),
0
0.2
2
Ttb
=1
(3 4R) = 0.5 + R,
(8.30)
Ttb
1.2
3

which is 0.5 for R = 0 (impulse turbine), but 0.833 for R = 0.5.

65

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

9
9.1

Lecture notes

Turbine solidity. Mass flow limits. Internal cooling.


Solidity and aerodynamic loading

Just as for the compressor, the blade spacing required to assure that the flow directions at the exit of the nozzles
and of the buckets are as intended, is determined by the aerodynamic loading limits of the blades. Instead of
the Diffusion Factor that we used to characterize this loading in the compressor, it is usual to define a Zweifel
Coefficient, which is a modified form of the blade lift coefficient
!
R x2
x
(p ps )d
x1 p
cx
z
,
(9.1)
0
ptb pc

Vb

vb=0

Vb

cx

c
Vc

Vc

vc =-r

r
Figure 9.1: Sketch of the flow across the rotor for the determination of the Zweifel coefficient. Note that in the
sketch vb0 = 0 while the equations in the text are for the case with vb0 6= 0
where x is the axial coordinate, pp is the pressure on the pressure side and ps is the pressure on the suction
side. This is a measure of the pressure difference across the blade divided by the outlet dynamic pressure. In
the ensuing argument we will relate it to the quantities that define the velocity triangle, just as we did for the
Diffusion Factor.
From a tangential momentum balance, if s is the blade spacing and c is the axial chord,
!
Z x2
x
0
0
(9.2)
c wc |vb vc | s = cx
(pp ps )d
cx
x1

66

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Substituting in the above definition of the Zweifel Coefficient and assuming that p0tc = p0tb , and Mc  1,



vb0
0
c wc s |vc | 1 0

vc
z =
(9.3)
1
02
c Vc cx
2
with

vc0
wc
0
=
cos

and
= sin c0 leading to
c
Vc0
Vc0
0


0




cx
0 vb
0
0 vb

z
= 2 sin c cos c 0 1 = sin(2c ) 0 1
s
va
vc

(9.4)

which equals sin(2c0 ) when vb0 = 0.


Zweifels design rule is that loss is minimized for 0.8 < z < 1.

9.2

Mass flow per unit of annulus area and blade stress

Since the pressure and hence density are higher at the inlet of the turbine than at the compressor inlet one
might think that the mass flow per unit area would not be an important factor in the turbine. Instead, it is
relevant because for a given blade speed the stresses at the root of the blade increase with blade length and
hence with the annulus area.
This is readily seen by computing the stress at the blade root. Consider a blade rotating as in figure 9.2.
The stress is given by
Z rT
2
r2 rH
2
(rH ) =
2 rdr = 2 T
= Af low
(9.5)
2
2
rH

Figure 9.2: A prismatic bar rotating about an axis through its end, illustrating the origin of centrifugal stress
in rotating parts of an engine.
This is often quoted as the AN 2 limit on turbines, N being the rpm. One can see that for a given blade
tip speed the stress increases as the radius of the root decreases, as is necessary to increase the flow area.
Returning to the question of the mass flow per unit area, we can compare stations a & b upstream and
downstream of the nozzle as in the situation shown in figure 8.1. Suppose for simplicity that the vanes have
zero thickness and that the radial height of the passages is constant. We can then deduce a connection between
b and Mb . Continuity of the axial flow requires that

a Ma Ta
a Va
.
a Va = b Vb cos b cos b =
=
(9.6)
b Vb
b Mb Tb

67

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

The temperatures and densities at the two stations are related by


1 2
1+
Ma
Tb
2
=
Ta
1 2
1+
Mb
2

b
=
a

Tb
Ta

1
 1

(9.7)

and substituting these we find


+1 )
1 2 2(1
1+
Mb
Ma
2

cos b =
(9.8)
Mb
1 2
1+
Ma
2
Suppose Mb = 1, then we see that as b increases and therefore cos b decreases, Ma decreases, so the mass
flow per unit area decreases. Recalling that we wanted large b in order to get large work from the turbine we
see that there is a tradeoff between high work and low stresses.

9.3

Turbine cooling. General trends and systems. Internal cooling.

As we have learned from our performance analyses for turbojets and turbofans, the thrust per unit of airflow
of the aircraft engines increases monotonically with turbine inlet temperature ratio, which we have designated
as t . Though it is less obvious, the thermal efficiency and specific impulse also increase due to the increase in
t . Thus, there is a powerful incentive to increase the turbine inlet temperature.
The trend of temperature increase with time is shown in figure 9.3.
In the early period, the increase was limited by metallurgical progress in developing oxidation-resistant
materials, primarily Nickel and Cobalt based alloys, that could operate for long times at high temperatures.
Beginning in the mid-60s, the technology of air cooling was introduced, and most advancement since then has
been due to more refined cooling techniques, although the introduction of directionally-solidified and singlecrystal blade materials has allowed some increase in metal temperature.
Cooling Systems: The general scheme for air cooling of the hot section of an engine is shown in figure 9.4.
In most cases the cooling air is drawn from the compressor discharge because it must be at a higher pressure
than that of the part of the flow path to be cooled. This leads to the important condition that the cooling
air temperature is the compressor discharge temperature. As shown in the diagram, some of the air is used to
cool the turbine nozzles, by a combination of internal convection cooling and film cooling, processes that will
be discussed in detail below. Another portion of the cooling air is transferred onto the rotor and used to cool
the rotating blades and their supporting disc. Here again a combination of internal convection and film cooling
is used.
Internal Convection Cooling:
In this form of cooling, the blade temperature is maintained below that of the hot gas passing through
the turbine by a flow of compressor-discharge air through passages internal to the blade, as suggested in the
diagram. To model this process we must therefore model first the heat transfer to the outside of the blade from
the hot gas, then the heat transfer on the inside to the coolant. The objective in cooling design usually is to
maintain the blade temperature nearly constant in the presence of a variation of the heat flux to its surface
from the hot gas.
External Heat Transfer:
Let us look first at the heat transfer to the outside of blade. We can write it as
qw = St [uCp (Tr Tw )]

(9.9)

where u is the mass flow density in the hot gas, Tr is the recovery temperature and Tw is the wall
temperature. The definition of Tr is the temperature the wall would reach if adiabatic, and it is close to the
68

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

As we have learned from our performance analyses for turbojets and turbofans, the thrust
per unit of airflow of the aircraft engines increases monotonically with turbine inlet
temperature ratio, which we have designated as t . Though it is less obvious, the thermal
efficiency and specific impulse also increase due to the increase in t . Thus, there is a
powerful incentive to increase the turbine inlet temperature.
Aerospace Propulsion

Lecture notes

The trend of temperature increase with time is shown in the figure:

Figurethe
9.3:increase
The trend
of limited
temperature
inlet temperature
with time
In the early period,
was
by metallurgical
progress
in developing
oxidation-resistant materials, primarily Nickel and Cobalt based alloys, that could operate
forstagnation
long times
at high temperatures.
in the
external
temperature.
St is defined Beginning
by this relation,
butmid-!"#$%&'()&')*(+,-,./&,0&123&
what is its physical significance? We have
cooling was introduced, and most advancement since then has been due to more refined
2cqw
2St ,
(9.10)
cooling techniques, although the introduction
of=directionally-solidified
and single-crystal
uCp (Tr Tw )s
blade materials has allowed some increase in metal temperature.
which is equal to the fraction of heat flowing through the blade row which enters the blade. Here as before c is
the blade
chord,Systems:
s is the blade
spacing, and = c/s is the solidity.
Cooling
4()&.)+)31-&$*()5)&0,3&123&*,,-2+.&,0&'()&6(,'&$)*'2,+7&,0&1+&)+.2+)&2$&
Fortunately, St is a small number because the boundary layers on the blades are relatively thick compared
shown in the figure below:
to the blade chord. In fact St depends on the details of the boundary layer behavior, and of course, is not
a constant, but for our purposes here we can think of St as being an average value for the blade surface. To
estimate its magnitude we can model the blades as flat plates, in which case the local St is given approximately
by:
2St
2St

0.66
= Cf =
Rex
0.0592
= Cf =
1/5
Rex

laminar

(9.11)

turbulent

(9.12)

where Rex is the Reynolds number based on distance from the leading edge of the plate. One important
difficulty for turbine cooling arises because transition from laminar to turbulent flow occurs at some Reynolds
number in the range of 3 105 to 106 . Looking at the expressions for St we see that for example if the transition
69

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 9.4: Schematic of air-cooled turbine, with cross section of cooled airfoil section at top. Note three modes
of cooling are present here: (a) Ordinary convection cooling by flow parallel to surface. (b) Impingement cooling
at the blades leading edge. (c) Film cooling over the blades suction side.
occurs at Rex = 106 , then at the transition point,
St

St

0.66 103 (laminar)


0.037 (turbulent)

(9.13)
(9.14)

So the heat transfer jumps suddenly by a large factor at the point of transition.
To get an idea of the magnitude of Rex in an engine let us look at some typical numbers.
For pt4 = 20 atm, t = 6, Tt = 6(298) = 1788 K, Rex = 8.3 107 /m at M = 1, so for a blade that has a
chord of 4 cm, Rex based on chord is 3.1 106 and we would expect transition on the blade surface. Actually
external flow turbulence tends to promote the transition so that it will probably occur at a lower value than
this. But the exact location of the transition is difficult to predict.
This gives rise to a major problem in scheduling the internal cooling system so that internal cooling matches
the external heat transfer.
Internal cooling rate:
For the internal cooling flow we model the flow as that through a long thin passage, for which
2 St |int = Cf |int = 0.023

1
0.2
ReD

(9.15)

One important question to be addressed is how much cooling air is required. Forming a heat balance such
that the heat flux into the blade equals that absorbed by the cooling flow,
2cqw = 2 St |e [uCp (Tr Tw )s] m
c Cp (Tw Tc )

(9.16)

where m
c is the cooling air mass flow rate per unit span, and Tc is the temperature at which the cooling air
enters the blade. Therefore, we obtain
Tr Tw
m
c
2 St |e
.
(9.17)
us
Tw Tc
70

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

A typical value of St |e = 0.005. For Tr Tw = 200 K, Tw Tc = 400 K, and = 1,


m
c
1
2(0.005) ' 0.005
us
2

(9.18)

Actually a ratio more nearly 0.01 is required, because the cooling air is not heated all the way to the blade
temperature. This is per blade row, not counting casing cooling. For a two-stage turbine, including these effects
we may need 2-4% of the air flow.

71

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion
16.50 L ecture 29

10
10.1

Lecture notes

Topics: F ilm cooling; T hermal stresses; Impingement

Film
cooling.
stresses. Impingement.
Cooling;
How to doThermal
cooling design
Film
Film cooling
cooling:

With internal cooling the heat flux through the blades surface increases as the external gas temperature
!"#$%"&#'(&)*%+,,*"&-%#$'%$')#%.*/0%#$(,/-$%#$'%1*)2'34%3/(.)+'%"&+(')3'3%)3%#$'%'0#'(&)*%
increases. That
is large Tr Tw implies large qw . This large heat flux must be conducted through the blade
gas temperature
increases. That is large Tr 5 Tw implies large qw. This large heat flux
material according
to
must be conducted through the blade materialT
according
to
w
(10.1)
Tqww = t
qw = k
t
where t is the thickness of blade skin.

where t is the thickness of blade skin.

So for a given back-side T, a small t is needed if the surface facing the hot gas is to be
10.1: Sketch of a thin skin with an applied heat flux qw
maintained atFigure
an acceptable
level. As noted in the last lecture, it is also necessary to
match the internal cooling precisely to external heat load to prevent a large Tw variation.
So for aAll
given
back-side
T , a small
is needed somewhat
if the surface
facing
the hot
gas isasto
be maintained at an
of these
difficulties
can bet mitigated
by the
technique
known
film
acceptable level.
As
noted
in
the
previous
section,
it
is
also
necessary
to
match
the
internal
cooling precisely to
cooling, which is illustrated in the figure:
external heat load to prevent a large Tw variation. All of these difficulties can be mitigated somewhat by the
technique known as film cooling, which is illustrated in figure 10.2.

Essentially the idea is to introduce cooling air into the boundary layer over the surface of
the blade, reducing the effective temperature of the gas heating the surface. This lowers
the heat flux, making less cooling air necessary. In addition the cooling air cools the wall
in passing through the small holes.
Figure 10.2: Schematic of cooling configuration for film cooling, showing a staggered double row of holes.

6$'%7'(.,(8)&+'%,.%."*8%+,,*"&-%"3%('*)#'2%#,%#$'%98)33%:'*,+"#;%()#",<

Essentially the idea is to introduce cooling air into the boundary layer over the surface of the blade, reducing
f
the effective temperature of thef ugas
heating the surface. This lowers the heat flux, making less cooling air
m=
necessary. In addition the cooling
air
cools the wall in passing through the small holes.
u
s s
The performance
of
film
cooling
is
related
film-to-stream
mass velocity ratio
where f refers to the film coolant
andtosthe
to the
free stream flow:
m=

f uf
s us

(10.2)

where f refers to the film coolant and s to the free stream flow.
As indicated in figure 10.3, which shows heat transfer rate versus the (dimensionless) magnitude of the
cooling air flow, the cooling air actually has two effects on the heat transfer to the wall, which can be separated
by varying the cooling air temperature relative to the wall temperature. If the cooling air is at the wall
72

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 10.3: Effects of film-cooling mass flow and cooling airflow temperature on heat flux to film-cooled flat
plate.
temperature its effect is to increase the heat transfer, by stirring the boundary layer. But if it is cooler than
the wall, this effect is partially offset by the reduction of the effective hot gas temperature, so that there is a
best cooling air flow rate.
To make these trends quantitative for design purposes, we define a Cooling Effectiveness by the relation
ad =

Tr Trf
Tr Tc

(10.3)

where Tr is the adiabatic recovery temperature, i.e. the temperature the wall would reach if adiabatic in
the absence of film cooling, Trf is the recovery temperature in the presence of film cooling, and Tc is the coolant
temperature, which is the lowest Trf . The cooling effectiveness so defined applies to a specific geometrical
arrangement of cooling holes, and also on the cooling mass flow rate. A set of data for a single row of holes is
shown in figure 10.4. The effectiveness decreases downstream of the row of holes, and in general increases with
cooling flow, at least for the range shown.
Now with this cooling effectiveness, how do we estimate the heat transfer rate to the wall? The usual
approach is to assume that the conventional film coefficient h, or the equivalent Stanton number St , can be
applied, with an effective gas temperature given by the cooling effectiveness, Thus,
h=

qw
Trf Tw

(10.4)

where h = uCp St , as usual, and


Trf Tw = Tr Tw ad (Tr Tc ).

(10.5)

qw = uCp (Trf Tw )St

(10.6)

We then calculate
where h or St are correlated in the usual way to the Reynolds number and Prandtl number.

73

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 10.4: Cooling effectiveness as a function of holes arrangement for various mass velocity ratios.

10.2

Impingement cooling

Because the heat transfer rate is largest at the leading edge of the blade where the boundary layer thickness is
least, special measures are sometimes called for to cool this region. One is Impingement Cooling, illustrated in
figure 10.5.
For the scheme shown the cooling rate can be correlated as
 0.4 !
 0.8
ld d
d
0.7
exp 0.85
(10.7)
N ustag = 0.44Red
p
dp L
where the quantities d, p, l and L are defined in Fig. 10.5 and the Reynoldss number is that for the impinging
jet
(u)jet d
Red =
(10.8)

and the Nusselt Number similarly is based on the jet diameter d


Nu =

10.3

hd
f luid

(10.9)

Thermal stresses

For a thin skin supporting a temperature difference due to a heat flux, as shown in Fig. 10.6, a thermal stress
is generated by the tendency of the hot side of the wall to expand relative to the colder side.
To see this we can note that in general the strain in the material due to a combination of a normal stress c
and a temperature difference from some mean temperature is
= (1 )

c
+ (T T ).
E

(10.10)

We denote with the sub-index 1 the cold side, i.e. T1 , 1 , etc, and with the sub-index 2 the hot side. The
mean temperature is T = 12 (T1 + T2 ). Note that the strain due to the temperature difference can be seen as

74

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

originated by a stress. In this approximation the stress, called the Thermal Stress, is found by equating the two
strains. For example, for the cold side
E
(10.11)
T 1 =
(T1 T )
1
If the plate is unloaded, force balance imposes H = c .
The temperature difference is related to the heat flux through the plate by
T1 T2 =

tqw
.

(10.12)

where is the thermal conductiviy of the material. Combining, we obtain


T1 =

E
qw t.
2(1 )

(10.13)

E
is termed the thermal stress susceptibility. Clearly we would like it to be small. It is
The quantity

proportional to the thermal expansion coefficient, and to the elastic modulus and inversely to the thermal
conductivity. Unfortunately the oxidation-resistant Nickel and Cobalt alloys favoured for turbine blades have
relatively low thermal conductivity and high modulus, so thermal stress is a controlling problem for turbine
blades. It results in cracking an ultimately limits the blade life.
(1 )U lt.
A related material quantity is the Figure of Merit, Z =
(in K), which is a measure of the
E
9
allowable temperature difference across the plate . A few values are given in Table 10.1.
Material
Z (K)

Cu
35

SS 302
111

Ti
136

Alloy Steel
234

Table 10.1: Figure of Merit Z for various materials.


The Z value is low for copper, but, of course, even a small temperature difference can drive a large heat flux
through copper which has high thermal conductivity.

10.4

How to design cooled blades

It is presumptuous to engage this subject at the tail end of a short lecture, because this is one of the most
difficult areas of gas turbine design. But the intent here is simply to outline the way one might proceed, as
follows:
(a) Estimate the qw over the blade surface for specified Tw .
(b) Find the thermal stresses in the skin.
(c) Find the reduction in qw required to limit the stresses to acceptable values, hence the ad required of the
film cooling.
(d) Find the arrangement of cooling holes and cooling air flow for film cooling, to give the required effectiveness.
(e) Find the internal cooling airflow to absorb the residual qw .

9 In

75

fact, using eqs. (10.12) and (10.13), we can see that Z =

1
(TH Te )max .
2

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

For the scheme shown the cooling rate can be correlated as


.8

l d d
0.85( )( )( )0.4
!
$
d
Nustag = 0.44 Re.7d # & e d p L
" p%

!"#$#%&"#%'#()*+,-.-%)/01#$%2-%&"3&%4*r the impinging jet


( u)applied
Jet d to the leading edge of a turbine blade. Radial exhaust
Figure 10.5: Schematics of Impingement
Cooling
Re
=
d
at left, chordwise exhaust at right.
and the Nusselt Number similarly is based on the jet diameter d
hd
Nu =
k fluid
Thermal stresses
For a thin skin supporting a temperature difference due to a heat flux,

a thermal stress is generated by the tendency of the hot side of the wall to expand relative
Figure 10.6: A thin skin supporting a temperature difference due to a heat flux.
to the colder side.
To see this we can note that in general the strain in the material due to
a combination of a normal stress !c and a temperature difference from some mean
temperature is

+ (T T )
E
76
Bioengineering and Aerospace Engineering Dept.
v. 2015-2016
If the plate extends to infinity in both directions, then the strains on both the hot and cold
sides will be the same (note however that this assumes the plate does not curve or bulge).
In this approximation the stress, called the Thermal Stress, is found by equating the two
strains:
= (1

Aerospace Propulsion

11

Lecture notes

Combustion: Combustors and Pollutants

11.1

Combustion process

The combustor is the heart of every aircraft engine. The combustor (and the afterburner, if any) is the only
component where we actually supply energy to the airflow, and we do this by burning a fuel that releases large
ammounts of heat.
Therefore, the combustion process is a central part of the operation of any engine. The time it requires
to completely burn the fuel since its injection is critical for the operation and sizing of the combustor. It is
necessary to complete the reaction before the fuel and air leave the combustor chamber and enter the turbine.
Some very important aspects of the combustion process that affect the combustion time are: (1) atomization
and vaporization (and their required times), (2) the actual fuel-to-air ratio and the stoichiometric ratio, (3) the
molecular dynamics that enable the combustion reaction to take place, (4) the mixing process and flame fronts
(diffusive, premixed, turbulent). These topics are discussed in the next paragraphs.
The fuels used commonly in aircraft engines (kerosene, JP-4), react with air only in gaseous form, so they
must be atomized, vaporized, and mixed thoroughly with the air, starting from liquid form. Most engines inject
the propellant into the combustion chamber as a spray (atomization), where the droplet size is determinant for
the required vaporization time (the smaller the droplet, the faster it will vaporize). Since it is important to
minimize the total time between fuel injection and reaction completion compared to the residence time of the
air in the combustion chamber, some engines also pre-vaporize the fuel before injecting it.
Even after full vaporization, the fuels burn only in nearly stoichiometric proportions (i.e., the ones dictated
by the reaction equation). However, the required fuel-to-air ratio in actual engines is far lower than the
stoichiometric ratio. As an example, take a stoichiometric mixture of octane (C3 H8 ) and oxygen:
C3 H8 + 5O2 3CO2 + 4H2 O
(11.1)

3(12) + 8
fuel
=
(0.23) = 0.0633
(11.2)
air stoich
5(32)
(here, 0.23 is the fraction of oxygen in air), whereas normal fuel/air ratios in jet engines are around 0.03. The
actual ratio of fuel/air over the stoichiometric ratio of fuel/air is called the fuel equivalence ratio, , and is here
seen to be as as low as 0.03/0.063=0.45. Normal hydrocarbon fuels do not ignite this far from stoichiometry:
even at the relatively high compressor discharge temperature, a value ' 0.9-1.1 is required. Hence, it
is indispensable that be carry out the combustion reaction in a primary region where near-stoichiometric
conditions are met; after the reaction, the excess air is then diluted into the reaction products to lower the
temperature of the mixture to the admisible conditions at the turbine.


Reaction
Products

fuel & air


mixture
vflame

!
!
!
!
!
!
"#$!%&'$(!)*!+%)+&,&'-).!)*!(/0#!*1&2$(!&%$!3$'$%2-.$3!45!&!4&1&.0$!4$'6$$.!
Figure 11.1: Flame front propagating into the premixed fuel and air.
%$&0'-).!'-2$!&.3!+%)3/0'(!3-**/(-).!'-2$7!&.3!&%$!%)/,#15!&(!(#)6.!-.!'#$!*-,/%$8!
! molecular mixture, the reaction takes place when the correct molecular collisions occur. There is a
In the

1
reaction activation energy, A, and the kinetic energy of impact between the reacting molecules mv 2 must exceed
2
77

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

A in order that a reaction to occur. Since the molecules have a kinetic energy according to MaxwellBoltzmann
distribution


mv 2
f (v 2 /2) exp
(11.3)
2Rg T


1
A
the fraction of molecules with mv 2 > A is exp
. Additionally, the number of collisions depends
2
Rg T
on the pressure of the reactants. As a result, the number of reacting collisions per unit volume and time, called
the reaction rate, is


A
m
reaction rate p f (T ) exp
(11.4)
RT
where m = 2 if 2 molecules are involved, 3 if 3 molecules are involved, etc. For typical hydrocarbons reacting
with air, m ' 2.
The spatial distribution of the reactants in the mixture is also important. When the fuel is not mixed, but
concentrated in droplets, only diffusion combustion can take place at the droplet boundary. This involves the
diffusive transport of fuel from the droplet into the combustion interface, and fresh oxygen from the exterior
into the droplet (similar to the combustion of a candle, where wax and air are diffused into the flame front
where the reaction takes place). This type of combustion is slower than when the propellant is well-mixed with
the air.
Combustion in premixed gases is the next simplest case to visualize and in a sense the easiest to experiment
on, although is not often the actual mechanism in practical combustion systems. We imagine a laminar flame
front propagating into the premixed fuel and air as shown in Fig. 11.1
The rates of propagation of such flames are determined by a balance between reaction time and products
diffusion time, and are roughly as shown in figure 11.2.

Figure 11.2: Laminar flame speed as a function of equivalence ratio for propane and hydrogen
In an actual engine combustor, however, the combustion process is fully turbulent. While turbulence is
responsible for part of the total pressure losses in the combustor, it is essential (and actually purposely enhanced)
to have a fast burning rate, since turbulence increases it with respect to the laminar one by at least two
mechanisms:

78

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

(a) wrinkling of the flame front: as the turbulence deforms the flame front, its area increases, and more reaction
per unit time can take place.
(b) stretching of the flame front: turbulence makes the flame front thinner, with the consequence of larger
gradients in concentrations, temperature, and pressure accross it. Since diffusion rates are dependent on
the gradients, this also increases the reaction rate.
Both of these mechanisms are promoted by a high level of fluid turbulence that is generated by the jets of air
entering the primary zone and the mixing zone.

11.2

Combustor chambers

The combustor of an aircraft engine releases an enormous amount of heat per unit volume, second only to highpressure rocket engines. In designing the combustor we must implement solutions to the following difficulties,
related to the combustion aspects described above.
(a) The temperatures reached during combustion are higher than the best oxidation-resistant materials can
withstand. This is overcome by protecting the combustion chamber walls by injecting a colder air sheet
through holes on the walls (see Figure 11.3).
(b) The overall mixture is very lean, and varies over a wide range with changes in the power setting of the engine.
As mentioned above, kerosene-air mixtures will only burn in a narrow range of mixtures near stoichiometric,
so the chamber is divided into a primary zone where near-stoichiometric ratios are achieved, and a dilution
region where the rest of the air is mixed with the combustion results. This is met by introducing the fuel as
a spray into a primary zone where the combustion is near stoichiometric and is stabilized by a recirculating
flow generated by the swirling motion to increase the residence time. The gas leaving the primary zone is
then cooled by mixing with additional air before it exits to the turbine nozzles.

Figure 11.3: Typical gas turbine combustor, showing primary zone of near-stoichiometric combustion followed
by dilution with excess air to reach Tt4 .
The combustor is typically constructed as a ring of can-like combustion chambers, each independent of each
other, or as a single annular combustion chamber (see Figure 11.4).
There are also a number of secondary requirements to be met by the combustor, some of which are quite
challenging too:

79

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 11.4: Combustors based on a ring of fire tubes or cans (left), and annular-combustion-chamber based
(right)
The burner must be dimensioned to assure re-starting at high altitude, when pressure, and hence chemical
rates, are low. Designs that are optimized for cruise may not meet this requirement.
The combustor must have an operating life of over 20,000 h., or some 5,000 cycles.
Pollutant formation must be below legally allowed limits (see par. 11.5).
The total pressure drop must be small, despite the strong flow turbulence that is purposely introduced
for enhanced mixing.
The temperature profile at combustor exit must be uniform (or sometimes must meet some prescribed
radial profile: lower at the root of the turbine blades, where the load is higher, and at the blade tips,
where cooling is more complicated).

11.3

Combustor sizing

The combustor must be large enough to allow the completion of the combustion to take place. The number
of processes that influence the combustion is so large that it is difficult to identify the driving requirement to
size the combustor. Since reaction rate increases with temperature and pressure, combustion efficiency tends to
decrease with decreasing entrance pressure and temperature. It decreases too with the entrance flow velocity,
since the residence time of the mixture in the burner, tres Ab 3 L/m,
decreases.
The chemical energy density in the burner is the mass of fuel per unit volume, times the fuel heat value,
E/V = a f h. If the time to full combustion, tburn is known, we can then calculate the power density as
P
a f h
=
V
tburn

(11.5)

In this section we will estimate the required dimensions of the combustor in the high-pressure (design
conditions) and low-pressure limits (high altitude and certain transcient operation). The burning time must be
estimated differently for each of this conditions.

80

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

(a) For design conditions the pressure and temperature are large enough, and the rate-limiting step is not
the reaction itself, but the turbulent mixing of the fuel and air, or of the rich primary burnt gas with the
dilution air. Let lt be a characteristic turbulent eddy size (of the order of the combustor diameter or width),
and ut some characteristic eddy velocity (some fraction of the mean flow velocity in the burner). Then we
estimate tburn lt /ut . For a numerical estimate, take a = 3 kg/m3 , f = 0.027, h = 43 MJ/kg, ut = 50
m/s and lt = 0.1 m. We then calculate a power density P/V = 1.7 109 W/m3 which is of the right order
of magnitude.
We can now calculate the combustor volume Vc = Pth /(P/V ), where the total thermal power is Pth = m
a f h,
and the airflow can be expressed as m
a = a ux Ac . Substituting and simplifying, the combustor length,
Lc = Vc /Ac is found to be
ux
lt
(11.6)
Lc
ut
which simply says that the combustor length scales with its width, or with whatever determines the eddy
size. We could call this photographic scaling.
(b) At low pressure, chemical reactions limit the combustion time, and assuming a bi-molecular reaction mechanism,
nproducts
nproducts
C(T )
tburn =
=

(11.7)
n products
R(T )n1 n2
a
The power per unit volume is now proportional to the air density, and repeating the steps of case (a), we
arrive at a combustor length estimate
ux
(11.8)
Lc C(T )
a
which is independent of width or any other size parameter. In other words, the chemically-limited combustor
length is independent of engine size.
Roughly speaking, the larger of the two estimates must be chosen in order to satisfy conditions both at
design and at low pressure. For a given engine family, the smaller engines will be chemically limited and will
all tend to have equal length combustors (so the combustor is most prominent in the smallest of them), but at
some larger size, length will need to be increased in proportion to width in order to obtain sufficient mixing.

11.4

Afterburners

The afterburner faces a different set of requirements:


(a) High outlet temperature
(b) High flow Mach number
(c) Low pressure drop required when non-afterburning
(d) Accepts flow from both core and fan
Like in the main combustor, the wall of the afterburner is slitted and colder airflow is allowed into the chamber
to create a protecting film for the walls.
In afterburners, the fuel is atomized at the exit of the turbine, to enable turbulent mixing. However, given
the large velocity of the flow, we need to implement some mechanism to hold the flame. This is because the
flame speed, even accounting for turbulent effects, is not as large as the mean flow speed, and the flame would
be carried away without some recirculatory zone where combustion has time to occur. Typically, the flame is
stabilized by bluff bodies, usually triangular in shape as indicated in figure 11.5.

81

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 11.5: Schematic of an afterburner, showing V-gutter flame holders stabilizing flame by producing a
recirculation zone (lower diagram)
The requirement for stability is that the fuel-air mixture spend a certain time, tchem , that is required for the
chemical reactions to initiate, in the mixing layer adjacent the recirculating wake. So the flow time tres must
be greater than this chemical time tchem . From the flow geometry,
tres =

L
ux

(11.9)

So we require

L
> tchem
(11.10)
ux
The
of the wake region is nearly independent of the absolute breadth, so we can say that
 length/breadth

L
L
L=
D, with
' 3 for normal conditions, and we see that a certain size of flame holder is required such
D
D
that
 
D L
> tchem
(11.11)
ux D
Since ux and tchem are independent of engine size, we find that D is as well. This is an example of lack of
geometric scaling, analogous to the low-pressure scaling for the main combustor.
Lastly, the length of the afterburner must be sufficient to allow all the fuel to be burnt. This, in practice,
requires that the turbulent flame fronts of each flame holder meet with each other and cover the whole flow
section. The turbulent flame front has a typical divergence half-angle of 3 deg. Naturally, the length of the
afterburner is reduced if more flame holders are employed; however, a trade-off exists on the number of flame
holders, since a larger number also incurs in larger pressure losses.

11.5

Pollutants: regulations

The motivations for control of emissions from aircraft, other than purely esthetic or ethical, stem from legislation
regulating the amounts of emission in the neighborhood of airports, and the possibility of regulation of emissions
82

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 11.6: (Left): Afterburner view of a MiG-23. Note also the actuators of the variable-geometry nozzle.
(Right): Mach disks in the high-Mach number plume of an afterburner.
into the upper atmosphere.
In 1973 in the US, the EPA published standards, to go into effect in 1979. They are compared in table 11.1
to the actual emissions from three engines in wide use at the time the standards were issued.
Pollutant

1979 Std

CO
U HC
N Ox
smoke

4.3
0.8
3
19-20

Actual
JT8-D
19
2.7
8
visible

emissions
JT9D-7
10.4
4.8
6.5
4

CF6-50
10.8
4.3
7.7
1.3

Table 11.1: Comparison of 1979 EPA Proposed Standards to Actual Emissions. All values are grams per kg
thrust, per hr. in standard approach-taxi-takeoff cycle. The smoke number, SN = 100(1 Rs /Rw ) is the
reduction in optical transmission through a glass witness plate exposed for a prescribed time, where Rs is the
reflectance of the sample and Rw is the reflectance of the clean paper. U HC stands for Unburnt Hydrocarbons.
Since the Federal Aviation Administration (FAA10 ) has had the regulatory jurisdiction over civil aviation, the
implementation of these standards fell to it. In July 1973 FAA issued SFAR27 to enforce only smoke regulations
but 1979 regulations were never incorporated into regulating process: manufacturers lobbied against them on
the basis that the increasing pressure ratios required for better fuel consumption would make their realization
difficult. In 1990 FAR 34 was issued by the FAA, regulating only smoke and unburned hydrocarbons, and
setting the limits for these essentially at the values attained by the oldest engines in operation.
In 1981, ICAO11 issued international emissions standards in terms of g/kg fuel, as displayed in table 11.2.
The Committee for Aviation Environmental Protection (CAEP), an international organization under ICAO,
has periodically updated the emission limits on N OX . The various CAEP regulations, including CAEP-6, in
effect as of 2008, are displayed in figure 11.7. The units on the ordinate are in grams of pollutant per kN of
thrust, averaged over a standard airport turnaround cycle. The abscissa is the overall pressure ratio of the
engine. For reference, the CAEP-1 regulation is from 1986, and the CAEP-4 is from 1998. It is clear that the
10 www.faa.gov
11 www.icao.int

83

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Pollutant
CO
U HC
N Ox
smoke

Standard
118
19.6
40 + 200
0.274
83.6F00

Table 11.2: 1981 ICAO international emissions standards in terms of g/kg fuel. 00 is the rated pressure ratio
and F00 the rated thrust in Newtons
trend is to tighten the emission rules over time, and also that commercial engines do comply, as they must,
with the rules in effect at the time they enter service (and usually ahead of time, in anticipation of further
strengthening of the rules).
In the European Union, The European Aviation Safety Agency (EASA12 ) is the institution in charge of the
regulation and certification of any air activities, devices, and people involved with them. EASA releases the
norms that have to be satisfied by any aircraft willing to fly in the EU airspace, and in particular, sets limits
on the emission of pollutants and noise and describes how to demonstrate compliance with them. These limits,
in turn, are those dictated by the CAEP of ICAO.
Since its beginning, EASA has striven to be intercompatible with FAA. To simplify the certification standards
(CS) produced by EASA, they follow the same numbering as the FAR from FAA: for example, FAR23 and
CS23 both contain the certification conditions for normal aircraft, and are essentially identical except for a few
exceptions.

11.6

Mechanisms for pollutant formation

The principal pollutants are Unburned Hydrocarbons, Carbon Monoxide, and Nitrogen Oxides. They are formed
as follows.
U HC - mostly at idle and low power, due to poor atomization etc. Also from venting of fuel systems.
CO - Due to fuel-rich combustion in primary zone.
Smoke - Due to fuel-rich droplet burning by diffusion flames
NOx - Following an element of fuel and air through the combustor we would see the following history of
temperature and concentrations.
During this transport, N OX is formed by the following sequence of reactions.
N2 + O
N + O2
N + OH

NO + N

(11.12)

NO + H

(11.14)

NO + O

(11.13)

where the concentrations of O and OH controlling the first and third reactions are determined by thermal
dissociation of O2 and H2 O. Taken together, the three reactions convert N2 to N O at a rate which is largely
controlled by the rate of the first reaction. The rate of N O formation may be represented as
d[N O]
= 2k[N2 ][O]
dt

(11.15)

12 www.easa.eu.int

84

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

operation.
The Committee for Aviation Environmental Protection (CAEP), an international organization
under ICAO, has periodically updated the emission limits on NOX. The various CAEP
regulations, including CAEP-6 , in effect as of 2008, are displayed in the figure below. The
units on the ordinate are in grams of pollutant per kN of thrust, averaged over a standard
airport turnaround cycle. The abscissa is the overall pressure ratio of the engine. For
reference, the CAEP-1 regulation is from 1986, and the CAEP-4 is from 1998. It is clear that
the trend is to tighten the emission rules over time, and also that commercial engines do Lecture notes
Aerospace Propulsion
comply, as they must, with the rules in effect at the time they enter service (and usually ahead
of time, in anticipation of further strengthening of the rules).

Figure 11.7: Historical evolution of emission limits on N OX as a function of overall pressure ratio, issued by
Mechanisms for formation of pollutants;
CAEP.

The principal pollutants are Unburned Hydrocarbons, Carbon Monoxide, and Nitrogen

where [ ] denotes
concentration
particles per unit volume of the indicated chemical species and k is a
Oxides.the
They
are formed asinfollows.
reaction-rate coefficient. According to the scheme outlined above, [O] is determined by thermal dissociation of
O2
O2 2O,
(11.16)
so that if the equilibrium constant for this reaction is
Kp (T ) =
we find

p2O
p O2

1 d[N O]
2k p
=
pO2 Kp .
[N2 ] dt
RT

(11.17)

(11.18)

This rate of formation is plotted as a function of temperature and equivalence ratio in figure 11.9 (for an initial
temperature of 700 K and pressure of 15 atm.).
Numerical example of N OX formation
From Fig. 11.9 the rate of N O formation in the primary (hot) zone of a kerosene burner at 15 atm is roughly
1 d[N O]
a
' 0.6 s1 . Let t be the residence time of the fuel in this zone, and use [N2 ] ' 0.77
(Wa = 28.9
[N2 ] dt
Wa
g/mol of air). Then
a
[N O] ' 0.6 0.77
t
(11.19)
Wa

85

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 11.8: Schematic time histories of temperature and species concentrations in an air-fuel mixture passing
through a combustor
is the concentration of N O generated. This is converted to mass density by multiplying times WN O , the
molecular mass of N O. The mean density of fuel in the mixture is f a , where f = m
f /m
a ' 0.066 at the
steichiometric conditions of the primary zone. Hence the mass of N O per unit of mass of fuel is
N O
a
1
0.6 0.77 30
t ' 7.3t.
= 0.6 0.77
tWN O
=
f
Wa
f a
0.006 28.9

(11.20)

For comparison with pollution standards, we convert this to (g N O)/(kg fuel) and obtain 7.3103 t(s). Compare
now to Table 11.2, assuming 00 = 15:
(N OX )allowed = 40 + 2 15 = 70g/(kg fuel).

(11.21)

This implies a limitation of the fuel residence time in the primary zone, such that
7.3 103 t < 70

t < 9.6ms.

(11.22)

In reality, this time covers a part of the time to mix with the secondary air, as reactions continue as long as
partially mixed hot pockets of mixture survive.
Strategies for reducing N OX
(a) Premix and burn lean. Nitrogen oxides are then formed at a low rate due to the low temperature, but
combustion tends to be unstable.
(b) Burn rich and then dilute and cool quickly. N OX is not formed during rich combustion, due to lack of
oxygen, but can form during mixing with air, so this part should be fast (hard to do).

11.7

Upper-Atmospheric Emissions

One of the main threats here is to the upper-atmosphere ozone, 03 . No regulations are in place yet for aircraft,
although there is an international agreement limiting the manufacture and sale of fluorocarbons, compounds
that severely damage the ozone layer.
86

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 11.9: Rate of change of N O mass fraction as a function of temperature and equivalence ratio, for
Cn H2n+2 initially at 700 K and 15 atm.
(a) Formation of 03 (the Chapman cycle):
Molecular oxygen provides a reservoir of O atoms through photodissociation by UV solar radiation:
O2 + h O + O

(11.23)

The atomic oxygen reacts rapidly in a three-body reaction with molecular oxygen, to form ozone:
O + O2 + M O3 + M

(11.24)

The population of ozone is mainly controlled by its destruction though photodissociation:


O3 + h O + O2

(11.25)

(b) Effect of Nitrogen oxides:


The natural effect is the continuous upwards transport of N O2 from the troposphere and its photodissociation to N O and O. This is followed by ozone destruction by
N O + O3 O2 + N O2
87

Bioengineering and Aerospace Engineering Dept.

(11.26)
v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 11.10: Variation of ozone concentration with altitude.


The natural transit time of N O2 from the troposphere is as long as 100 years, so even a small amount of
N O and N O2 released in-situ by high-flying aircraft could potentially lead to a strong acceleration of the
ozone destruction.
(c) Other effects:
In addition to these N OX concerns, recent information suggests that the particulate content of the exhaust,
largely sulfuric acid drops formed from the sulfur in the fuel, is a major factor in the overall effect on the
atmosphere, as these drops act as condensation nuclei leading to cloud formation.
Besides N OX interfering with the ozone creation-destruction process, engine operation is associated to
other detrimental effects on the upper atmosphere. However, we have only begun to understand them, and
their consequences are still a topic of active research.
The released CO2 from the combustion process is a well-known green house gas. Although presently the
emissions of CO2 due to air transport account for only 2% of the human-generated CO2 , Air transport is an
economic activity on the rise, with a 5% annual increase. Attention is being paid to reducing CO2 emissions
from aircraft engines.
The hot jet released by the engines triggers the condensation of the air humidity and the formation of clouds
(contrails, or linear clouds on the trail of the aircraft, and cirrus). Some of these clouds can last for several
hours or even days. While clouds per se are inoffensive, they increase the albedo (reflectivity) of our planet,
effectively decreasing the amount of sun radiation reaching the ground. The continued creation of artificial
clouds can therefore have an impact on the climate of the planet.

88

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

12

Lecture notes

Introduction to engine noise and aeroacoustics

Noise generation in the neighborhoods of busy airports has been a serious problem since the advent of the
jet-powered transport, in the late 1950s. Although piston-engined aircraft had caused some concerns prior to
that, it was the turbojet-powered first generation jets (707, Comet, DC-8) with their jet noise that led to wide
public concern, and to regulations by some airport authorities. With the continuing growth of airline travel,
the problem has continued to expand, leading to rules promulgated by the FAA that limit the noise that any
individual aircraft can make at each of 3 measuring stations.

Figure 12.1: Schematic of airport runway showing approach, takeoff, and sideline noise-measurement stations.
Likewise, EASA established the European norms for aircraft noise (based on those of ICAO).
Noise is the human ears response to pressure fluctuations in time, at the ear of the observer.13 Noise is
transmitted from the source (aircraft) to the observer by sound waves. So let us begin with a brief review of
sound propagation.

12.1

Noise propagation

Conservation of mass
Conservation of momentum

D
+ u = 0
Dt

(12.1)

Du
= p
Dt

(12.2)

Suppose the velocity and pressure to be the sum of a large steady component and a small time-varying one
u = u0 + u0 (t)
p

(12.3)

= p0 + p (t)

(12.4)

= 0 + 0 (t)

(12.5)

and further suppose that u0 = 0. Then to first order in small quantities eqns. (12.1) and (12.2) become
0
+ 0 u 0
t
u0
0
t

(12.6)

= p0

(12.7)
0

p
quantitative measure of noise is the decibel (dB), defined from the actual pressure fluctuation p0 (in Pa) as 20 log10 210
5 .
For reference, a pressure fluctuation p0 = 1 Pa (' 105 atm) yields 20 log10 21015 = 94dB. This is a loud noise already, so rather
small pressure perturbations can become a problem.
13 The

89

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

In the absence of heat conduction and viscosity, 0 and p0 are related isentropically, i.e.
tiating

dp
d
p0
0
=

=
p0
0
p0
0

It follows that

p0 =

p0 0
= RT0 0 a20 0
0

p0
p
= . Differen
0
(12.8)
(12.9)

where a0 is the speed of sound. Since a0 does not depend on time, we can write
1 p0
0
= 2
t
a0 t

(12.10)

Now substituting eq. (12.10) into eq. (12.6), differentiating with respect to time, and substituting eq. (12.7),
we obtain an equation for p0
2 p0
= a20 2 p0
(12.11)
t2
In one dimension (for a plane wave)
2 0
2 p0
2 p
=
a
(12.12)
0
t2
x2
This is a Wave Equation, satisfied in general by solutions of the form p0 = p01 (x a0 t) + p02 (x + a0 t), so that
solutions for p0 that are constant on x = a0 t. Thus the solution for a plane wave would be of the form
or

12.2

p0 = P cos(x a0 t)

(12.13)

p0 = P cos(x + a0 t)

(12.14)

Acoustic energy density and power flux

Let us form the combination [Eq. (12.6)]

p0
+ [Eq. (12.7)]u0 = 0, leading to
0

u0
p0 p0
+ p0 u0 + 0 u0
+ u0 p0 = 0
2
0 a0 t
t
which can be re-written as

1
p02
0 |u0 |2 +
2
20 a20

+ (p0 u0 ) = 0.

(12.15)

(12.16)

1
The first group in parentheses has dimensions of Energy/Volume, and the part 0 |u0 |2 is easily identified as
2
the kinetic energy per unit volume due to acoustic oscillations. The second term is the compression potential
E
energy per unit volume due to the pressure and density fluctuations. Thus, the first part of eq. 12.16 is
,
t
where
1
p02
E = 0 |u0 |2 +
(12.17)
2
20 a20


E
is the total energy density. Its rate of decrease
must be due to the divergence of the acoustic power
t
per unit volume, and so we define the acoustic power flux as the vector
s = p0 u0 .
90

Bioengineering and Aerospace Engineering Dept.

(12.18)
v. 2015-2016

Aerospace Propulsion

12.3

Lecture notes

Noise sources and noise modeling

Any source of fluctuating pressures gives rise to noise. Aeroacoustics is the study of noise generation and
propagation by the interaction of the fluid with moving surfaces and due to turbulence. We can now ask how
such waves are generated in an engine. Some of the main sound sources are sketched in Fig. 12.2:

Figure 12.2: Sources of unsteady flow giving rise to noise from aircraft engines
Elementary sources. A general noise source can be decomposed as the superposition of different elementary
multipole sources. Each type has different radiation characteristics. It is useful to examine the simplest acoustic
sources, which are all configurations with an imposed pressure fluctuation or an imposed wall vibration. The
first in a systematic series of such sources are shown in Fig. 12.3.
(a) Monopoles
Suppose now the perturbation is due to a pulsating sphere, whose radius oscillates by some small r0 at a
frequency . This sets up spherically symmetric pressure field that oscillates at as well, and propagates
as a sound wave train. In spherical coordinates, the wave equation is


0
2 p0
a20
2 p
r
.
(12.19)
= 2
t2
r r
r
We expect acoustic energy to be conserved, and since at least the compression part of this energy varies as
p02 , let us try a solution of the Monopole form:
r 
0
p0 = P
cos [k(r a0 t)] .
(12.20)
r
91

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 12.3: Monopole, dipole, and quadrupole sources and some fluid disturbances they can represent.
It can be seen by direct substitution that this does satisfy the wave equation. The quantity k =
2
a0
= 2
.
k

The velocity field (purely radial) can be calculated from equation (12.7)

is called
a0

the wave number, and the wavelength is =

u0r
p0
=
t
r

and using eq. (12.20), after integrating in time,




sin [k(r a0 t)]
P r0
cos [k(r a0 t)]
u0r =
0 a0 r
kr

(12.21)

(12.22)

The acoustic power flux (energy crossing unit area per unit time, averaged over one cycle) is the average of
eq.(12.18), where u0 is purely radial:
Z
1 0 0
P 2  r0 2
=
p ur dt =
(12.23)
0
20 a0 r
92

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

where =

Lecture notes

2
2
=
. The net acoustic radiated power is

ka0
Pm = 4r2 = 2

P 2 r02
0 a0

(12.24)

which is seen to be independent of r, as it should. The monopole power radiated is independent of wave
2
, and only dependent on pressure amplitude. A possible physical implementation of a
number k =

monopole source is a pulsating jet, such as produced by a pulse-jet (like that in the V-1 missile), or by the
oscillations during an engine surge.
(b) Dipoles
Consider next two monopoles of equal strength P operating in counter-phase to each other, and spaced a
small distance d along the xdirection. If an observer is located at a distance r from one of them, and at
an angle from the xdirection, its distance to the other will be (approximately) r + d cos . Then the
pressure p0 at the observers location will vary as
p0 = P

r0
r0
cos [k(r a0 t)] P
cos [k(r + d cos a0 t)]
r
r + d cos

(12.25)

Expanding the second term and assuming d to be much smaller than both the wavelength (2/k) and the
observation distance r,
p0 = P

r0
r0 d
kd cos sin [k(r a0 t)] + P
cos cos [k(r a0 t)] .
r
r r

(12.26)

Of the two terms in eq. (12.26), the first has the 1/r dependence that will ensure energy flux conservation
at all distances, while the second will decay faster and will be negligible for distances r  1/k; this second
term is the near-field dipole sound, which can be important near the source. We concentrate here on the
far-field contribution (first term in eq. (12.26)).
The calculation of the far-field power flux is as before, remembering that it is still a radial flux, even though
there is an angular dependence in . So we still have the first equality in eq. (12.23), and the calculation
is straightforward. We obtain
2
1  r0
(r, ) =
P kd cos
(12.27)
20 a0
r
cos2
distribution (no sound at 90 to the dipole axis, maximum along the dipole
r2
axis). The total radiated power is
Z
2 (P r0 kd)2
Pd = 2
(r, )r2 sin d =
(12.28)
3
0 a 0
0
This flux has now a

Aside from the 1/3 factor, this differs from Pm , the monopole power, by the factor (kd)2 = (2d/)2 , which
is, by assumption a small number, and becomes smaller the longer the wavelength is (or the lower the
frequency = k/a0 ). This is, of course, due to partial cancellation of the fields of the dipoles.
An important observation is that a physical dipole requires application of an external oscillatory force. A
possible physical implementation of the radiating dipole is any vibrating compact object, such as a fan or a
turbine blade. The dipole axis is then the direction of vibratory motion, and the surrounding air is forced
back and forth as it would between the hypothetical two monopoles in counter-phase.

93

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

(c) Quadrupoles
Strongly turbulent flows, such as an engine exhaust jet, are known to be strong sources of acoustic radiation.
If the jet is steady and subsonic, there is no possibility of macroscopic monopole (expansion/contraction)
type of radiation, and since there is no external force in the body of the fluid, no dipole sources either.
However, there are fluctuating pressures at different points (turbulent eddies), exerting forces on each other
with zero net force on the larger scale. The lowest order multipole with these features is the Quadrupole,
which can be built up from two dipoles with a common axis, and separated by 2d and with opposing
directions.
The detailed derivation is similar to that for a dipole. We calculate (in the far field)
r0
p0 = 2P (kd cos )2 cos [k(r a0 t)]
r
and for the acoustic power flux,
2  r0 2
P
(kd cos )4
(r, ) =
0 a0
r
which integrates for all directions to a radiated power
Pq =

8 (P r0 )2 (kd)4

5
0 a0

(12.29)

(12.30)

(12.31)

The quadrupole radiator pattern is more sharply directional along the axis (cos4 ), and is an additional
(kd)2 weaker than the dipole, with an even stronger rate of increase with frequency.
It should be noted that the collinear-dipole type of quadrupole is not the only one possible, but all of them
share the (kd)4 feature (although with different angular patterns).

12.4

Jet Noise

Having reviewed the mechanisms for noise propagation and the three basic types of acoustic sources Monopoles,
Dipoles and Quadrupoles, we can now examine aircraft engines as sources of noise. We begin with jet noise
because historically it was the principal source of concern with the introduction of jet transports. Jet noise is
generated by the turbulent mixing of the high speed exhaust jet with the ambient air. We can think of the
mixing region as a region of strong turbulent flow, as shown in Fig. 12.4, that can be modeled as a random
array of Quadrupoles, since there is no mass source, and no force in the fluid.
In a turbulent jet, in order of magnitude p 0 u2 (u being the jet velocity), and the size of the fluctuating
regions is some fraction of the diameter D of the jet. Similarly, the spacing d of these regions must also be some

factor times D. The wavenumber k =


can be estimated by noting that u/D. We then have,
a0
2 
4
0 u2 D
u
u8 D2
Pq
D 0 5 .
(12.32)
0 a0
Da0
a0
The striking feature is the scaling of this radiated power with the 8th power of the jet velocity, a result due
to Lighthill (1963). To compare it to the jet kinetic power, note that
Pjet (0 uD2 )u2 = 0 u3 D2
and therefore
Pq

Pjet

u
a0

5

5
= Mjet

(12.33)

(12.34)

This very strong sensitivity to jet velocity (or Mach number) is one of the significant disadvantages of the pure
turbojet as compared to the high bypass turbofan (the other being its lower propulsive efficiency, although that
disappears at a high enough flight Mach number).
94

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 12.4: A subsonic jet mixing with ambient air, showing the mixing layer followed by the fully developed
jet.

12.5

Turbomachinery noise

Vibrations of lifting blades, or periodic passage of lifting blades past the observer, produce dipole-type radiation,

,
which radiates power as given by eq. (12.28). Estimating in this case r0 d c (blade chord), k
a0
a0 c
where u is now the blade speed, and p = 0 u2 , the power radiated per blade is
Pbl

0 u6 c2
a30

(12.35)

The force on one blade is of the order of Fbl 0 u2 cH 0 u2 c2 , and so the acoustic efficiency is
 3
Pbl
u
bl.noise =

= MT3 .
Fbl u
a0

(12.36)

This is noise generated, but the duct, even with no special absorbing design, attenuates the sound, and in
fact, forbids propagation for the lower range of frequencies. The frequency is here related to the blade Mach
number MT and the number B of blades
Brot rT
rT
=
= BMT
a0
a0

(12.37)

and so MT must be higher than a certain threshold for blade noise to effectively propagate.
For a source moving at subsonic speed, the sound is in fact exponentially attenuated over a distance of the
s
order of 2
. For a supersonic source the waves from the infinity of pressure peaks and troughs on the wall
interfere constructively with each other. For a subsonic source there is no such direction, and interference is
always destructive.
In the case of turbomachinery, the supersonic relative velocity can come from two sources
(a) Supersonic tip speed
(b) Subsonic tip speed, but interacting blade rows
To understand the second situation, consider a rotor operating in the vicinity of a stator. The stator blades
produce a pressure variation
p0stator (incidence) exp(iV )
(12.38)
95

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

where V is the number of stationary vanes.


The disturbances from the rotor are periodic in rotor coordinates, so
incidence exp[iB( R t)]

(12.39)

where B is the number of moving blades. It follows that


p0stator exp [i{(V B) BR t}]
So the Effective Speed of Rotation of the disturbance is

d
BR
ef f =
=

dt const.phase
V B

(12.40)

(12.41)

In early engines there was a tendency to make V B and they were very noisy. Now we use the rule
V = 2B + 1, so that the rotor-stator interaction is no more likely to generate propagating disturbances than is
the rotor rotation itself.

96

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

13

Lecture notes

Engine rotating structures

Rotating structures are a crucial component of any modern engine. Centrifugal forces on a differencial mass
dm located at a radius r and rotating at a rotational speed are expressed as dF .
dF = 2 rdm = 2 r2 drddz

(13.1)

It is thus intuitive that centrifugal stresses can be extremely relevant due tu the high rotational speeds of
compressor and turbine rotors. In the design phase, we want to ensure that the rotating parts can withstand
the tremendous rotational velocities; that the hot parts can survive the high thermal stresses; and that the
whole engine can support the pressure and thrust forces. At the same time, we want to minimize the mass of
the system and simplify its construction assembly.

13.1

Blade loads

The centrifugal stresses on the blades are straightforward if we assume them constant over the section:
Z rT
A(r) =
2 A(r)r dr.
(13.2)
r

An exponential decrease is thus required for = const through the rotor. Moreover, the gas creates an axial
and a tangential bending moment on the blades. In this regard, the blades behave as cantilever beams. The
differential forces that give rise to these moments are here expressed for the case of a compressor rotor assuming
constant axial velocity:
BdFz = 2r(pb pa )drBdF = gas w2r(vb va )dr

(13.3)

Bending moments are maximum at the hub and the full blade mass is supported by the root stress. These
stresses have to be taken into account when designing the blades and rotor disk. In particular, considering only
the centrifugal stresses:
Z
b 2 rT
rAb (r)dr.
(13.4)
bH =
Abh rH

13.2

Centrifugal stresses and disc design

By disc we refer to the rotating structural members that carry the rotating blades in the turbomachine. They
are unusual as structural members in that they must withstand very large tensile stresses generated by the
centrifugal forces. Their mass is a large part of the total mass of an engine, and they set a limit on blade speed,
so an understanding of their characteristics is essential to appreciate the performance limits of modern aircraft
engines.
The rotor disk is ended in an outer rim, which carries the blades, and an inner rim. By design, in order to
utilize the material most efficiently, we wish to keep the stress level constant throughout the disc. We begin
by describing the stress state of a small volume element of a disc with the usual elasticity analysis. Taking to
be isotropic in and r, the net radial force on the element is given by


dz
d
z 2 r2 ddr = zrd (r + dr) z +
d + 2zdr
(13.5)
dr
2

97

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 13.1: Schematic of a turbomachine rotor, consisting of disc (with the inner and outer rims) and blades,
with free-body diagram showing stresses
The force balance is then
z 2 r2 ddr
1 dz
z dr
z
so that

dz
drd,
dr
2 r
=
,


= r

zI
= exp
z0

(13.6)
(13.7)

const exp


2 (r02 rI2 )
2

2 r
2


2
,

(13.8)


.

(13.9)

For the outer rim, consider the radial force balance, as in the bottom left sketch. There are B blades, not
necessarily at the same stress as the disc, and we shall assume their centrifugal pull is smoothly distributed
around the periphery
0 W0 T0 (rH d) 2 rH +

bH Ab B
d = zH rH d + W0 T0 d.
2

(13.10)

In equation (13.10), the first term corresponds to the rim centrifugal force, the second to the centrifugal pull
from blades, the third to the support from rim stress and the last one to the radial force from hoop stress. We
obtain
B
Ab
2
0 2 rH
+ bH
2
W0 T0
=
.
(13.11)
zH rH
1+
W0 T0
98

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

So the disc supports the rim, and reduces its stress. Now we add an inner rim; it supports the disc and
there are no blade loads, so in this case
I WI TI (rI d) 2 rI
I 2 rI2
=
.
zI rI
1
WI TI

= zI rI d + WI TI d,

(13.12)
(13.13)

With this simple stress model, we can establish some general guidelines for the design of a disk.
(a) From aerodynamics, we know or choose rT , rH /rT , B, Ab (r), W0
(b) Set permissible stresses and bH (rT may then be limited by eq. (13.4))
(c) Choose a T0 and get zH from eq. (13.11)
(d) Choose an rI and get zI from eq. (13.9)
(e) Get WI TI from eq. (13.13)
This leaves T0 and rI as free parameters to be selected by other considerations.

99

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

!
!
"#$!%&'()*+'!+%!)#$!,-&$$.$!%*/0,!1*//!2$!3*,(&,,$3!*'!)#$!'$4)!/$()&5$6!!!
! Figure 14.1: Cross section of typical main-shaft ball and roller bearings
753*'85*/9!$8(#!5+)8)*':!,;++/!*,!,&;;+5)$3!29!+'$!28//!2$85*':!)#8)!;+,*)*+',!*)!
84*8//9!8'3!8/,+!82,+52,!583*8/!/+83,<!8'3!+'$!+5!0+5$!5+//$5!2$85*':,!)#8)!8(($;)!
14 Fundamentals
of rotordynamics
583*8/!/+83,!2&)!8//+1!84*8/!0+=$0$')!)+!8((+00+38)$!)#$508/!$4;8',*+'!8'3!
,)5&()&58/!3$%+508)*+',6!
Vibrations of the rotating mass in the engine are critical aspect of their operation. The engine rotor should be
!
designed to be stiff and well balanced, although no shaft is ever 100% balanced. Hence, provisions must exist
"1+>,;++/!,&;;+5)!8558':$0$'),!85$!,#+1'!,(#$08)*(8//9!*'!)#$!%*:&5$?!
to avoid entering resonances and to damp out the vibrations without allowing their growth beyond limit.
!
Vibrations induce additional, time-varying loads on the bearings of the engine which must be taken into
account when designing these. The blade tip clearance must be sufficient to avoid impacting the engine casing
even in the largest amplitude oscillations.
Oscillations appear not only on the rotating elements of the engine, but on every part of its structure. The
study of vibrations in the engine is a complex subject, since a realistic analysis must include a many-degrees-offreedom model, and account for the time-dependent interaction between the solid parts and the stream, which
can excite or damp certain modes of vibration.

14.1

Bearings and engine arrangements

As noted, engines are unusual amongst engineering structures in that such a large fraction of their total mass is
rotating at high speeds. This large rotating mass must be supported on bearings so as to maintain quite close
clearances between the blade tips and the stationary casings, on the order of 1 mm on a rotor of 1 m diameter,
or one part in 103 . At the same time of course the stationary structure must be as light as possible.
All existing engines use ball and roller bearings to support the rotating assemblies, called spools. These
are shown schematically in figure 14.1.
The function of the squeeze oil films will be discussed in the next paragraph.
!
Ordinarily each
! rotating spool is supported by one ball bearing that positions it axially and also absorbs
radial loads, and one
or more roller bearings that accept radial loads but allow axial movement to accommodate
"#$!)+;!8558':$0$')!&,$,!@&,)!)1+!2$85*':,<!+'$!28//!8'3!)#$!+)#$5!5+//$5<!)+!
thermal expansion,&;;+5)!$8(#!+%!)#$!)1+!,;++/,6!!A)!*,!(+0;8()!8'3!8)!/$8,)!*'!;5*'(*;/$!5$/8)*=$/9!
and structural deformations.
Two-spool support
arrangements are shown schematically in figure 14.2.
/*:#)6!!"#$!/+1$5!8558':$0$')!&,$,!)#5$$!2$85*':,!+'!)#$!*''$5!,;++/<!+'$!28//!
The top arrangement
uses just two bearings, one ball and the other roller, to support each of the two spools.
8'3!)1+!5+//$5<!8'3!%+&5!2$85*':,!+'!)#$!+&)$5!B/+1!,;$$3C!,;++/<!+'$!28//!8'3!
It is compact and at least in principle relatively light. The lower arrangement uses three bearings on the inner
spool, one ball and two roller, and four bearings on the outer (low speed) spool, one ball and three roller.
Both of these arrangements have advantages and both have been used successfully in high-bypass commercial
transport engines. This is just one more illustration of the fact that design solutions are not unique.

100

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 14.2: Arrangements of two high-bypass turbofans (PW-JT9D (top) and GE CF-6 (bottom)).

14.2

Lumped mass model

In this section we derive a simplified model of the shaft vibrations, assuming that all its mass is concentrated
on a single point.
As noted in the last section, gas turbine engines are unusual structures in that a large fraction of their mass
is rotating at high speed. Thus, vibration and dynamics have always been a serious issue with them. What is
intended here is that the physical phenomena of principal importance are conveyed by study of a set of simplified
models.
Schematically, we can think of the rotating system as a mass on a flexible shaft, mounted on flexible supports
(Fig. 14.3). If most of the mass resides in the rotor and the part of the shaft adjacent to it, we can combine the
two flexibilities as two springs in series with a combined stiffness which tends to restore the rotors center to
the axis from whatever direction it is deflected into. More complex models, with additional degrees of freedom,
would be needed if the bearings and the parts of the shaft near them had a significant mass. We still retain the
simpler formulation, but will allow the rotating mass to be mounted slightly eccentrically, to study the loads
this can generate.
Fig. 14.4 shows the situation. The center of mass C has coordinates (x, y), and the geometrical shaft
center, also assumed to be its elastic center, is at S. The center of mass is offset by a distance e from S, and
rotates with the shaft at the angular rate . There is an elastic restoring force kOS directed from S to O,
as well as a viscous damping force, normally provided by an oil squeeze film damper between the shaft and
the bearing. This damping force is proportional to the rate of change of that part of OS that comes from the
bearing displacement, but since we ignore the mass of the bearings, that distance is a fixed fraction of OS itself
dOS
(depending on the ratio of shaft to bearing stiffness), and we model it as b
.
dt
We can now write a pair of equations of motion, one for x, the other for y
Mx
=
M y =

101

kxs bx s ,

kys by s ,

Bioengineering and Aerospace Engineering Dept.

(14.1)
(14.2)

v. 2015-2016

Aerospace Propulsion

Lecture notes

Figure 14.3: Model of a turbomachine rotor mounted on a flexible shaft mounted on flexible bearings.
where
xs
ys

= x e cos(t),
= y e sin(t).

(14.3)
(14.4)

It is actually more convenient to work with a complex displacement z = x+iy, so that a single complex equation
is needed:
M z =

kzs bzs ,

(14.5)

z e exp(it).
(14.6)
r
k
b
and the friction parameter n =
, and eliminating
Introducing the undamped natural frequency n =
M
M
zs , we obtain
z + nz + n2 z = e(n2 + in) exp(it).
(14.7)
zs

The general solution to this equation consists of the homogeneous part, with two arbitrary constants determined by initial conditions, plus the forced, or particular solution, determined by the right hand side. The
homogeneous solution is a superposition of damped sines and cosines at the natural frequency n . Because
of the damping, it will disappear some time after a transient event, leaving only the forced solution, which is
proportional to exp(it), like the right hand side of the equation. We concentrate here on the forced solution
only (the steady state). Putting z = B exp(it) and substituting into the governing equation, we can solve for
B, and find
z
zs

n2 + in
exp(it),
n2 2 + in
2
exp(it).
= e 2
n 2 + in

= e

(14.8)
(14.9)

The force on the supports is, in complex form,


F = kzs + bzs = M z = +M 2 z,
102

Bioengineering and Aerospace Engineering Dept.

(14.10)

v. 2015-2016

Aerospace Propulsion

Lecture notes

Y
t
x

C
e
y

Figure 14.4: Sketch of the model with the rotating mass mounted eccentrically.
where the last form can be directly rationalized as the result of centrifugating the center of mass. It is, of
course, a rotating force (around O), proportional to exp(it). The amplitude of this force is obtained from the
expression for z
s
n4 + (n)2
2
|F | = M e
(14.11)
2
(n 2 )2 + (n)2
and the phase can also be written down, although we omit it here.

14.3

Critical speed

The solutions for the amplitude of z (normalized by the eccentricity e) or for F (normalized by M 2 e) are
shown in Fig. 14.5.
Some important points are
1. With no damping, z/e and |F |/e for n (resonance condition, critical speed).
2. Damping limits the resonant amplitudes to finite values. The critical speed is slightly lower than n in
that case.
3. The system can be run above n with a flexible shaft or support, in which case S rotates about C for
very high speed (x and y tend to zero).

14.4

Forces on bearings

Now let us ask how the bearing stiffness and damping influence the forces on the structure.
(a) At subcritical speeds, namely, 2  n2 , F M 2 e exp(it) and the forces increase as the square of the
speed, as could be expected. Notice this condition might not mean a very low rotational speed, if the shaft
and the bearings are stiff enough, or the rotor is light enough to produce a very high natural frequency (this
is unusual in jet engines, though).
103

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

Aerospace Propulsion

Lecture notes

15

10

A
5

0
0

0.5

1.5

2.5

Figure 14.5: Amplification factor as a function of the rotating frequency. Black line, n/n = 0. Red line,
n/n = 0.1. Blue line, n/n = 0.2. Green line, n/n = 0.3.
n2
0, i.e., the center of
2
mass tends to remain stationary, with the shaft rotating about it. The force then tends to |F | M n2 e = ke,
a constant, rotating elastic force towards the center of mass that transmits to the supports and causes
vibrations. This is the operating condition for supercritical speed engines.

(b) At supercritical speeds, 2  n2 , and with relatively weak damping, |z| e

Elaborate procedures are followed to reduce eccentricities to a minimum, and the shaft can be periodically
re-balanced but with engine wear, the eccentricities are difficult to maintain near zero. Their effects can be
mitigated by (a) Staying away from operation at a speed equal to the natural (critical) frequency. (b) If
normal operation is at supercritical speed, crossing the critical range as quickly as possible, and (c) Providing
as much damping as practical with squeeze-film or other dampers.

14.5

Comments on blade vibrations

While the study of aeroelastic phenomena is complex, some general observations are gathered below:
1. Blades can oscillate independently, or cooperatively withing a given disk or with neighbouring disks
2. Bending and torsion modes exist
3. Due to the large density of the blades compared to the density of the air, the frequencies and modes are
essentially those of the blade in vacuum. The air stream does, however, provide excitation or damping to
these modes
4. Temperature changes the material modulus, and so the natural frequency of the blades
5. Blades passing each other and the overpressures that appear at that instant can be a major source for
excitation.
6. Blade flutter, i.e., the bending resonance with the air stream, typically requires that the blade be at or
near stall. However, collective modes exist for blade cascades where the interaction between blades and
flow passages is highly non-trivial.

104

Bioengineering and Aerospace Engineering Dept.

v. 2015-2016

You might also like