You are on page 1of 279

AXIAL FLOW FANS AND COMPRESSORS

To my grandson, Neil, without whose interest this book


might never have been completed

Cranfield Series on Turbomachinery Technology


Series Editor: Robin L. Elder
Axial Flow Fans and
Compressors
Aerodynamic Design and Performance

a. b. McKe n z ie
Visiting Professor
School o f Mechanical Engineering
Cranfield University
Bedford

Ashgate
Aldershot • Brookfield USA • Singapore • Sydney
O A. B. McKenzie 1997

All rights reserved. No pan of this publication may be reproduced, stored in a retrieval
system, or transmitted in any form or by any means, electronic, mechanical, photocopying,
recording or otherwise without the prior permission of the publisher.

Published by
Ashgate Publishing Limited
Gower House
Croft Road
Aldershot
Hants GU11 3HR
England

Ashgate Publishing Company


Old Post Road
Brookfield
Vermont 05036
USA

British Library Cataloguing in Publication Data

McKenzie, A. B.
Axial flow fans and compressors : aerodynamic design and
performance. - (Cranfield series in turbomachinery)
1. Axial flow compressors 2. Axial flow compressors - Design
3. Aerodynamics
I. Title
621.5' 1

Library of Congress Cataloging-in-Publkation Data

McKenzie, A. B., 1926-


Axial flow fans and compressors : aerodynamic design and
performance / A. B. McKenzie,
p. cm.
Includes bibliographical references.
ISBN 0-291-39850-2
1. Fans (Machinery)*-Aerodynamics. 2. Axial flow compressors-
-Aerodynamics. I. Title
TJ960.M35 1997 97-1249
621.6’ l-dc21 CIP

ISBN 0 29139 850 2


Printed in Great Britain by The Ipswich Book Company, Suffolk
Contents

Preface xi

1 The axial fan


Introduction 1
The domestic ventilator 2
Work done 3
The diffusing blade passage 4
Mass flow variation 7
The fan characteristic 9

2 Reaction
Introduction 12
Definition 13
Choice of reaction 15
Variation of reaction with radius 16

3 The work and flow coefficient diagram


Introduction 17
Additions to the diagram 18
Reactions other than 50% 22
The IGV/rotor fan 24
The rotor/stator fan 24

4 Blade loading parameters


Introduction 27
Cascade data 27
Total pressure losses 29
Deviation 29
Axial velocity ratio 30
Compressor data 30
Blade geometry selection 32
vi Axialflow fans and compressors

5 Blade geometry
Introduction 35
The C series of aerofoils 35
American practice 36
Cascade geometry 36
The double circular arc blade 37
Modem trends 38

6 Aspect ratio
Introduction 41
Surge margin improvement 41
Other benefits 42
Technical considerations 43
Equivalent cone angle 45
End wall loss parameter 45
Design criteria 47

7 Relative motion effects


Introduction 49
Blade design with secondary losses 50
Wakes in relative motion 52
Work and flow coefficient effects 53
Radial flow shift through the rotor 54

8 Vortex flow
Introduction 56
Free vortex designs 56
Constant reaction 57
Arbitrary vortex 60
Simple equilibrium 60
Examples of simple equilibrium 61
Constant ‘reaction* design 63
Throughflow computations 65

9 Mach number effects


Introduction 66
High subsonic Mach number • 66
Passage throat width 67
Supersonic Mach number 69
Blade forms for supersonic operation 70
Passage shock waves 70
S/C requirement 71
Laser anemometry 72
Effect of Mach number on overall pressure ratio 73
Contents

10 Reynolds number effects


Introduction
Performance variation
WasselTs correlation
Surface finish effects

11 Compressible flow relationships


Introduction
Notation
Fundamental relationships
Applications

12 The stage characteristic


Introduction
The work coefficient characteristic
The efficiency characteristic
The pressure rise coefficient
Experimental stage characteristics

13 The repeating stage concept


Introduction
The linear repeating stage
Stage of unity work coefficient

14 Spanwise matching of high diameter ratio stages


Introduction
Variable axial profile
Variable work done design
An alternative approach

15 Spanwise matching of low diameter ratio stages


Introduction
Example
Increased diameter ratio
Conclusions from the approximate method
Alternative methods

16 Stall
Introduction
Criteria for progressive or abrupt stall
Surge initiation
Rotating stall
Hysteresis
viii Axialflow fans and compressors

Vibration 119
Passive stall control 120
17 Surge
Introduction 124
An alternative criterion for surge 126
More complex approaches 129
Surge margin 129

18 Performance presentation
Introduction 132
Performance graphs 133
Isentropic and polytropic efficiency 136
Polytropic efficiency 138
Efficiency assessment 140
Efficiency with interstage bleed flows 142

19 Stage matching and surge control


Introduction 144
The stage matching problem 144
Bleed off-take 147
Variable stagger stator blades 149
Industrial axial compressors 150
Twin spool gas turbines 151
Off-design operation of twin spool gas turbines 153
Active surge control 156

20 Inlet flow maldistribution


Introduction 157
Compressors in parallel 158
Distortion parameter 161
Inlet Temperature distortion 162

21 Compressor rig testing


Introduction 164
Power requirements 164
Facility layout 168
Test procedure 171

22 Stage performance prediction


Introduction 174
A preferred method 174
Other methods 179
Off design efficiency 180
Contents ix

23 Overall performance prediction


Introduction 181
Stage stacking 181
Overall performance correlation 183
High speed corrections 188
Test comparison with prediction 190

24 Performance with altered gas properties


Introduction 191
The significance of density ratio 192
Gases other than air 193
Compression of helium 194
Viscosity and Reynolds number 195

25 Design for a domestic ventilator


Specification 196
Overall parameters 196
Blade speed and dimensions 197
Blade selection 198
Half speed operation 200

26 Design for an industrial fan


Specification 201
Preliminary considerations 201
Choice of diameter ratio 202
Rotor blade geometry 203
The outlet stator 206
The outlet diffuser 208

27 Design for a transonic wind tunnel fan


Specification 209
Preliminary considerations 209
Selection of vector geometry 210
The rotor blade 212
The stator 216
Annulus dimensions 218
Design speed 220
Fan efficiency 220
Outlet diffuser 223
Power requirement 224
Operation at higher tunnel Mach numbers 224

28 Design for an industrial compressor


Introduction 226
Design specification 226
X Axialflow fans and compressors

General considerations 227


Design parameter selection 227
Rotor blade geometry 230
Stator blade geometry 234
Annulus diagram 236
Variation of stage numbers 237
Additional front stages 238
Scaled versions 240
Volume flow 240

29 Outline design for a jet engine compressor


Introduction 241
Specification 241
Overall parameters 242
First stage 243
Last stage 245
Blading 247

30 HP compressor mid-stage
Introduction 248
Datum stage 248
Blade end modifications 249
Rotor design 251
Stator blade 253
Design method 255

31 The high bypass fan


Introduction 256
General arrangement 256
Aerodynamic parameters 257
Rotor blade geometry 260
Fan blade integrity 260

Bibliography 262
Preface

My introduction to axial compressors was at Rolls Royce in 1947 when the


technology was still in its infancy, so it was relatively easy to absorb existing
knowledge and thereafter keep abreast as the technology developed. The task
facing newcomers to the subject today must be vastly more formidable, faced
with literally thousands of technical papers produced over the last five decades
and growing at an ever greater pace.
Despite the wealth of detailed publications available there have been only a
few books which attempt a general introduction to the subject in sufficient depth
to provide an adequate understanding on which to base further research,
development and particularly aerodynamic design. It is my hope that this volume
will go some way to serve this purpose, and if only a few readers become as
fascinated by the subject as I have been for nearly fifty years, the effort will be
well rewarded.
A few novel concepts which may not be generally accepted have been
introduced. I feel these have sufficient merit to justify inclusion, if only to
stimulate further thought, which may confirm or reject them. Examples are to be
found in Chapters 7, 14, 15 and 17. A method of overall performance prediction
described in Chapter 23 also contains some features not previously published.
As is to be expected I have drawn extensively on my thirty seven years
experience with Rolls Royce, and I am grateful to the Company for permission
to publish some of the subject matter. This is not to say that Rolls Royce endorse
any of the material, the responsibility for which is entirely my own. I also wish
to thank Mr D P Hope of Rolls Royce for many helpful suggestions and much of
the material for Chapter 31.
I am also indebted to Professor R.L.Elder and Dr.K.Ramsden at Cranfield
University for their help and guidance in the preparation of the text.
1 The axial fan

Introduction

The simplest form of axial compressor is the single stage fan. The distinction
between a fan and a compressor has been defined by describing the fan as an air
mover, whereas the purpose of a compressor is to increase the pressure and
density of the air or other working gas. While this definition is substantially true,
nevertheless the specification of a fan usually calls for a pressure rise, even if
this is only a centimetre or two of water gauge. In turbofan aero-engines the fan
may develop a pressure ratio in the region of 2 : 1, but is still an air mover in that
the majority of the airflow passes directly to a propulsion nozzle where the
pressure energy is converted directly to kinetic energy in the propulsive je t

Figure 1.1 Axial ventilator fan


2 Axialflow fans and compressors

The domestic ventilator

A familiar form of axial flow fan is the domestic ventilator. This consists of a
single row of blades driven by an electric motor. An example is shown
diagrammatically in figure 1.1. Rotation of the blades induces a draft of air from
the room to the exterior. Examination of the section of the blades will show a
shape as in figure 1.2a or 1.2b. For the simplest designs the blade is of constant
thickness (called “cambered plate") but more sophisticated designs have an
aerofoil section i.e. similar to the cross section of an aircraft wing.

Fig. 1.2a Cambered plate blades

Fig. 1.2b

Figure 13 Rotor inlet and outlet vectors


The axial fan 3

When the fan is running the air flows towards the blades with an axial velocity
Va. Due to the blade motion, U, the velocity of the air relative to the moving
blade at its leading edge is V|, as shown in figure 1.3. The angle, ot|, of the
relative velocity from the axial direction is similar to the angle of the blade at its
leading edge, so that the flow passes smoothly along both blade surfaces. Due to
the curvature of the blade section, and the influence of neighbouring blades, the
flow relative to the blade is turned towards the axial direction as it passes
through the blade passage. The resulting vector diagram at the trailing edge is
also shown in figure 1.3. Due to the flow turning produced within the blade
passage the outlet flow has a circumferential velocity component, Vw3, in the
same direction as the blade movement It is assumed for the present that the axial
velocity does not change in passing through the blades.

Work done

Work is done when a force moves its point of application. The motor produces a
circumferential force on the blades and the work done by the movement of this
force is transmitted to the airflow.
m
The work done per second = Ft.U, and so x.co = F,.U

where Ft = tangential force on the blade


U = blade velocity = co.r
co = angular velocity of blades
r = radius of rotation
x = torque = F,.r

The torque acting on the rotor blades is reacted by the rate of change of
angular momentum of the air. Hence torque can be written:
t = MrVw3

where M is the rate of mass flow and Vw3 is the tangential velocity at rotor exit.

Hence: T od = MUVw3

The energy transfer to the airflow per unit mass flow is:
AH = tcd/M = UVw3
In more general terms this can be written as:
AH = U3VW3 - UqVwq
4 Axialflow fans and compressors

In this form allowance is made for the situation where the flow may leave the
rotor at a different radius from which it enters, and the blade speed changes from
Uo to U3. There may also be a whirl velocity ahead of the rotor, Vw0. In accurate
analysis and design work it is necessary to take account of radial changes across
the blade row. For immediate purposes it will be sufficient to write:

AH = U.AVw

where AVw is the change of whirl velocity across the rotor. It is convenient to
superimpose the inlet and outlet vector diagrams on a base of the common blade
speed as in fig. 1.4. This form of the velocity triangle diagram is fundamental to
an understanding of the design and performance of axial fans and compressors.

Fig. 1.4 Inlet and outlet vector diagrams superimposed

The diffusing blade passage

The relative inlet velocity, Vt, is reduced to V2 in passing through the blade row.
The inlet flow area of the relative flow is given by:

n s h cos a t

where n is the number of blades, s is the circumferential distance between blades


(the pitch or space) and h is the span or radial height of the blade. At blade outlet
the relative flow area becomes n.s.h.cos CX2. Continuity of mass flow gives:

M = p 1A] V1 = piAiVi

where p is the density of the air and A is the cross sectional flow area. For low
speeds compared to the speed of sound the density change can be assumed
negligible and hence:

V2/V 1 = Ai/A2 = cos <Xi/cos 012


The axialfan 5

Bernoulli’s equation for energy for incompressible flow, i.e. constant density,
can be written:

P = p, +V4pV,J = p2 + l/ipV22

where P is the total or stagnation pressure relative to the blades, which is


constant through the blade passage in the absence of losses, and p is the static
pressure. There is therefore a rise of static pressure through the blade passage
given by:

P2 - pi = V£pVj2- !^pV22

Dividing through by VipVj2 we can write:

Cpj = ApMpVj2 = 1 - (V2/VO2 = 1 - (cos ot|/cos (X2)2

The term V£pV]2 is the inlet dynamic pressure, equal to the total minus the
static pressure(i.e. P) - pt) where Pi is measured relative to the moving blade at
inlet. Experience has shown that Cpi should not be much greater than 0.5 for
efficient, stall free operation. This corresponds to a velocity ratio V2/V1 of 0.707.
Thus the amount of turning in the blade passage should be limited so as to keep
the outlet relative velocity greater than 70 per cent of the inlet relative velocity.
The rise of static pressure across the blade is accompanied by a loss of total
pressure relative to the blade, caused by friction of the air with the blade surfaces
and the annulus walls. This reduces the static pressure rise:

Pi = pi + V*pV,2 = p2 + V$pV22 + APl

where APl = Pi - P2. In the stationary frame of reference, however, the total
pressure increases:

P3 = p3 + V6pV32

and p3 = P2, since static quantities are the same in both frames of reference.

P3-Po = P3 + *pVj 2 -po-V4pV0J


6 Axialflow fans and compressors

AP = Ap + VipVj* - V4pV0J

For a low pressure rise fan such as a domestic ventilator the whirl angle 013 is
small, say less than 20° and V3 is not much greater than V0, hence:

AP = Ap (approximately)

For higher pressure rise fans a 3 may be 30° or 40° and V3 is significantly
greater than V0. In this case it becomes useful to fit a row of outlet guide vanes
(OGVs) in order to remove the whirl velocity and further increase the static
pressure, although some small loss of total pressure is inevitably involved. This
will be less than the dynamic pressure equivalent of the whirl velocity which
would be entirely lost if no guide vanes were fitted. The outlet guide vanes
provide an increment of static pressure rise from the air deflection:

p4 - P3 = '/Spv32 - VipV,2 - APl

where P4 and V4 are the static pressure and velocity downstream of the OGV and
APl is the loss of total pressure across the OGV. Note that when the OGV turns
the flow to the axial direction:

P« - Po = p« + VSpVa2 - po - fcpVa2

Hence: AP = Ap across the complete stage.

For low speed fans, which includes the majority of industrial machines, the
velocities involved are low enough in relation to the speed of sound that changes
of density in the flow can be ignored. The ratio of the air velocity to the local
speed of sound is called the Mach number, and provided this is less than 0.3
throughout the flow the density changes are less than 5%. At normal room
temperature this means that flow at velocities up to 100 m/s can be treated as
incompressible i.e. of constant density, for most engineering purposes. Where
the Mach number significantly exceeds 0.3 the flow must be treated as
compressible, and methods to be described in a later Chapter are necessary. The
fans used in modem turbofan aero engines operate with axial Mach numbers up
to 0.7, and the Mach number relative to the rotor blade tip at inlet is typically
about 1.5. Consideration of these fans is left until a later chapter. Meanwhile it
must be emphasised that while, for accurate calculation, it is necessary to treat
flows of greater than 100 m/s as compressible, it is convenient to use the
assumption of constant density for illustrative purposes, even for Mach numbers
up to say 0.6. It should also be noted that a great deal of compressor research
work is carried out on low speed machines at Mach numbers in the region of 0.2.
The axialfan 7

These machines are often quite large, having casing diameters up to 1.5 m. This
allows operation at Reynolds numbers of 2*1Or or more which avoids the
majority of the adverse effects on performance of small scale or low speed.
(Reynolds number is discussed in Chapter 10.)

Mass flow variation

If the mass flow through the fan is reduced from the design value the continuity
equation indicates that the axial velocity will fall in proportion:

M » pAVa

where M is mass flow, p is the air density, A is the flow area, and Va is the axial
velocity. The reduced axial velocity at a constant blade speed produces an
increase of the inlet angle relative to the moving rotor blade. The incidence is the
difference between the blade angle at its leading edge, pi, and the relative air
angle ai:

i = at - Pi

Thus the incidence is increased as axial velocity is reduced. The air direction
relative to the rotor at the trailing edge will increase only slightly for an increase
of incidence in the region of the design value. The outlet triangle will therefore
have almost the same a* at the reduced axial velocity, and the change of whirl
velocity across the rotor will be increased, i.e. the work input per unit mass flow
will be greater.
The increase of total pressure across the rotor, in the absence of losses can be
written:

P, - Po = P3 + *pV3J - Pi - ’/SpVa2

After some manipulation this can be reduced to:

APi/p = UVw3 = AH
8 Axialflow fans and compressors

where AP| is the ideal total pressure rise, i.e. without losses. The ratio of the
actual to ideal total pressure rise is the rotor efficiency, and so finally, we can
write:

AP/p = r|AH

where i\ is the efficiency. These quantities are most conveniently non-


dimensionalised by dividing by the blade speed squared to give:

AP/(pU2) = t|AH/U2

The left hand side of this expression is known as the total pressure rise
coefficient, and AH/U2 is the work coefficient. The difference between the ideal
and actual total pressure rise is the total pressure loss occurring in the rotor

APl = pAH - AP.

It can readily be shown that APl is also equal to the difference in the total
pressures relative to the moving rotor at inlet and outlet:

APl = P ,- P 2

It is this loss which is measured by stationary cascade tests and it is usually


expressed non dimensionally by dividing by the inlet dynamic pressure relative
to the blade (i.e. VipVj2). This is known as the total pressure loss coefficient and
is written as:

0 = APl/(H p V,j)

Curves of GJ against the incidence can be obtained from cascade tests. The inlet
and outlet total pressures are normally measured only at the midspan of the
cascade. The losses measured are therefore only those due to the blade profile
and do not include the end wall losses or the secondary losses associated with
flow curvature and wall boundary layers. The cascade midspan loss is therefore
called the profile loss coefficient, GJP. A typical plot of GJp against the incidence,
l, is shown in figure 1.5. This is for an inlet Mach number less than 0.3. As the
Mach number is increased beyond this the losses tend to increase, as will be
The axialfan 9

described in a later chapter. It is evident that the profile losses, as shown on


figure 1.5, rise rapidly when the incidence departs more than about ±10° from
the angle giving the minimum loss. For positive incidence this is due to
pronounced separation of the flow from the blade convex or suction surface. For
large negative incidence there can be separation from the concave or pressure
surface, but the flow can at these inlet angles reach sonic velocities in the blade
passage throat and can become ‘choked’. For these reasons, where the loss rises
shaxply at positive incidence is called ‘stall’, and at negative incidence ‘choke’.

Incktone*

Figure 1.5 Loss coefficient variation with incidence

The fan characteristic

The performance of a fan is generally expressed as the pressure rise v mass flow
characteristic, and these quantities are often conveniently expressed as the total
pressure rise coefficient, AP/(pU2), and the flow coefficient, Va/U. The work
coefficient, and the efficiency are also commonly plotted against Va/U, as in
figure 1.6 and 1.7. Assuming initially that the air outlet angle relative to the rotor
remains constant as the incidence is varied it is apparent from the vector diagram
that
AH/U2= 1 -ta n a 2Va/U

For a given constant value of a 2 this gives the straight line characteristic of
figure 1.6. The outlet angle in fact increases slowly with increasing incidence
until stall, when it may rise more rapidly than the incidence, thus causing a
maximum deflection to be reached in the blade passage. The true work
10 Axialflow fans and compressors

coefficient curve is therefore of slightly reducing slope as flow is reduced, and is


also shown on fig.1.6 .

now Coefficient

Figure 1.6 The work characteristic

Typically the efficiency curve has a shape as shown in figure 1.7. Since
efficiency can be written as:
i\ = 1 - APl/(pAH)

it is clear that the maximum efficiency will occur where AIV(pAH) is a


minimum. The work, AH, rises continuously as flow is reduced. It follows that if
the loss were constant with flow, the efficiency would also rise with the work.
Maximum efficiency therefore occurs not at minimum loss, but at a somewhat
lower flow, where the loss is rising in the same proportion as the work. The only
exception to this is when the gradient of the work v flow coefficient curve has a
zero or positive gradient, which is rarely the case. The curve of total pressure
rise coefficient v flow coefficient tends to peak as the efficiency falls and the
slope of the work curve is flattening. This is generally taken as the stage stall
point as operation at lower flows, if at all possible, is likely to be rough and
possibly mechanically damaging. Some fans designed at low work and flow
coefficients, say less than 0.2 and 0.3 respectively, may not exhibit a positive
gradient on the pressure rise curve, but merely a slight kink before continuing
with a negative gradient to zero flow.
The axialfan

Flow Co«ff.

Figure 1.7 Pressure rise and efficiency characteristics


2 Reaction

Introduction

The term derives from the reaction force created by a device such as a water
sprinkler where a rotor is driven by the reaction to the momentum of a jet
emerging tangendally to the direction of rotation. This is a pure reaction turbine.
Some early steam turbines were pure impulse, which is the opposite in that the
fixed nozzles give a high tangential velocity in the direction of rotation.

Fig. 2.1 Impulse turbine with near zero outlet whirl


Reaction 13

The rotors remove this tangential velocity and convert it to mechanical torque.
The outlet velocity in the stationary frame of reference may have little or no
tangential velocity as indicated in figure 2.1. There is no static pressure drop in
the rotor of such a turbine, and the relative inlet velocity equals the relative
outlet velocity, only changing direction. A pure reaction compressor would have
all the stage static pressure rise in the rotor. This is called 100% reaction. The
stator is impulse, i.e. it has no change of velocity or static pressure, only a
change of direction as shown in figure 2.2. A pure impulse compressor would
have no static pressure rise in the rotor, and the vector diagram is as in figure
2 .2 , but with the rotor and stator interchanged.
These extremes are seldom used in multi-stage compressors where the reaction
is generally within the range 50% to 90%. A form of single stage fan used for
special purposes does employ reaction in excess of 100%. An inlet guide vane of
high deflection, e.g. 45°, giving an inlet whirl velocity against rotation is
employed. The vector diagram is shown in figure 2.3 and some features of this
configuration are discussed in Chapter 3.

Rotor outlet Stator outlet


vector vector
Stator inlet
vector
Blade speed • n R0t0r inlet vector

Fig 2.2 100% reaction compressor stage

Figure 2 3 IGV and rotor stage of greater than 100% reaction

Definition

In multi-stage compressors the air direction leaving successive stages may vary,
as also may the axial velocity. For this reason it would be more convenient to
define the reaction as the static pressure change in the rotor divided by the total
pressure change in the stage. In pressure terms a knowledge of the losses is
14 Axialflow fans and compressors

required, it is simpler therefore to define reaction in terms of temperature


changes. This allows reaction to be expressed in terms of the vector diagram
quantities. The energy equation in terms of temperatures and velocities is:

T = t + Vl/2Cp

where T is the total, or stagnation, temperature, t is the static temperature, and Cp


is the specific heat at constant pressure. The static temperature rise in the rotor
is:

ti - ti * (T* - Vjl/2C,) - (T-. - V.V2C,)

The total temperature relative to the rotor, T ^ is constant through the blade
passage if the flow is at constant radius, so:

t, - 1, = (V ,2- V j2)/2Cp

The stage total temperature rise occurs entirely in the rotor for all reactions and
is derived by equating the enthalpy rise to the work:

CpAT = UAVw

Hence reaction, R = (t2 - t,)/AT = (V,J - V22)/(2U.AVw)

Assuming the axial velocity constant through the stage:

V,2 - V22 = Vw,2 - Vw22 = AVw(Vw, +Vw2)

Hence R = (Vw, +Vw2)/2U

For Vw, +Vw2 = U; R = 50%


For Vw, +Vw2 = 0 ; R = 0%
For Vw, +Vw2 = 2U; R = 100%

Reaction can also be expressed as:

R = 0.5(tana, + tana2)Va/U

and tancxTn= 0.5(tanai + tana2) by definition, hence:

R = tana„,Va/U

where a,,, refers to the rotor vector mean flow angle. Other useful relationships
are:
Reaction 15

Vw0 = U(1 - R) - AH/2U

Vw3 = U(1 - R) + AH/2U

Figure 2.4 indicates the derivation of these relationships.

AH/2U

^Vwo
I Jjl
^ RU
-__ Vw2_____ 'yJS R)U
Fig^.4 Reaction relationships

Choice of reaction

It can be argued that 50% reaction will give the best efficiency on the following
basis. Assume a fixed minimum loss coefficient for rotor and stator, irrespective
of reaction. At 50% reaction the inlet relative velocities to rotor and stator are
equal. Hence equal loss coefficients means equal losses in both blades. At any
other reaction the relative inlet velocities are unequal. Since the loss coefficient
is proportional to the square of the inlet velocity the blade with the increased
inlet velocity will have a greater increase of loss than the reduction on the other
blade. Thus there will be a reduction of efficiency as reaction departs in either
direction from 50%. On the other hand as reaction departs from 50% the blade
with the greater relative velocity at inlet has a reduced deflection and the other
an increased deflection. This may cause a reduction of loss coefficient where
deflection is reduced and vice versa, thus leaving the efficiency little changed. If
the inlet relative Mach number is raised above the critical on one blade by
departing from 50% reaction this will certainly have an adverse effect on the loss
coefficient and hence the efficiency.
Other considerations enter the decision at the first and last stages. For the fust
stage zero whirl at entry to the rotor avoids the need for a row of inlet guide
vanes to produce pre-whirl. This represents a reduction of aerodynamic loss, and
benefits of cost, weight, and length. The I.G.Vs. may require provision for anti­
icing which represents further saving of complexity and loss of engine cycle
16 Axial flow fans and compressors

efficiency in providing a hot air supply. Thus there has been a trend towards the
zero inlet whirl type of stage even when this means transonic relative inlet Mach
number onto the rotor with a consequent increase of shock losses.
For the rear stage of a compressor it is normal to remove all whirl velocity at
the exit This may require a stator row of high camber and low space chord ratio.
Some designers prefer a two row stator system. The losses in either case tend to
be high. By using a high reaction for the rear stages the deflection required in the
outlet stator is reduced and excessive losses are avoided. With arguments in
favour of near zero inlet whirl for first and last stages it is not surprising that this
style of stage design has been adopted throughout the compressor by some
designers with satisfactory results. The reaction of such stages rises as the work
coefficient falls, and is typically in the range 0.7 to 0.9.
For strictly zero inlet whirl the reaction is given by:

R = 1 - 0.5AH/U*

Variation of reaction with radius

There are a variety of ways in which reaction can be varied with radius. These
give rise to different forms of vortex flow and will be discussed under that
heading. It is worth noting here that the lower the reaction the greater is the mean
whirl velocity in relation to the blade speed. To satisfy radial equilibrium a
greater radial gradient of static pressure is therefore required, and this can lead to
an extreme gradient of axial velocity, high at the hub and low at the casing. This,
together with the extremely high deflections required for the inlet and outlet
guide vanes, has led to reactions of less than 50% seldom being employed.
3 The work and flow
coefficient diagram

Introduction

The diagram of work coefficient, AH/U2, versus flow coefficient, Va/U, as


developed in figure 1.6 for a single stage fan is of general use for design and
analysis of compressors. Instead of basing it on a stage with zero whirl velocity
at inlet to the rotor, it is more often drawn in terms of 50% reaction vector
diagrams, such as figure 3.1, where:

V,= V,; V ,-V ,; a, = a 3;anda0= a J

Figure 3.1 Typical 50% reaction vector diagram

AH = UAVw = U(U-2V w2)


18 Axialflow fans and compressors

AH/U1= 1 - (2VWj/U) = 1 - 2CVa/U)tanotj

This defines a family of straight lines radiating from VaAJ - 0, AH/U = 1.0.
Each line corresponds to a fixed value of Oj as indicated on figure 3.2. Note that
the lines become steeper as the value of increases. For a given blade geometry
the value of is approximately constant for a range of values of a, around the
optimum value. Variation of Oj at constant corresponds to a variation of axial
velocity, Va, at a constant blade speed, U, and therefore to a variation of the
flow coefficient VaAJ. Thus each of the lines is a first approximation to the work
v flow characteristic of a given blade geometry in a form independent of the
rotational speed.
O a- 0°

VaAJ

Figure 3.2 Lines of constant cti on AH/U2 v Va/U field

Additions to the diagram

Another set of straight lines can be defined in terms of a j instead of These


radiate from Va/U = 0, AH/U2 = -1.0. They are most easily located by noting that
<Xj = <Xj when AH/U = 0, and that at AH/U = 1.0, the flow coefficient, Va/U =
1/tana, The intersections of the two sets of lines give values of the air deflection
( e s O j - ctj) which takes place across the blade row. By joining points of equal
deflection the curves of figure 3.3 are obtained.
The work andflow coefficient diagram 19

Va/U

Figure 33 Lines of constant inlet angle and deflection

Yet another set of straight lines which are of interest radiate from the origin.
These must obviously be given b y :

AH/U1 = constant.Va/U

and from the 50% reaction vector diagram it is readily shown that:

U/Va = (Vwt + VWjVVa

U/Va = tan a, + tan Oj

By substituting for Va/U we can obtain:

(AH/U2XU/Va) = tana, - tano^


which can be written:

AH/(UVa) = tana, - tana^

This relationship can be obtained more directly from:

AH = UAVw

Hence AH/U = V w ,-V w 2

and AH/UVa = (Vw, - Vw2)/Va


20 Axialflow fans and compressors

or AH/UVa = tana, - tana^

Thus the straight lines radiating from the origin each represent a constant value
of difference of tangents of the inlet and oudet angles. It is also useful at this
point to define the vector mean angle of flow, amas:

tana^ = 0 ^(tana, + tana^)

on- tana 01= OJU/Va

which can be written as:

VaAJ = l/(2 tan aj

Curves for constant values of the de Haller number = Vj/V, can also be plotted
2
on the AHAJ v VaAJ field. Note that the de Haller number is directly related to
the ideal static pressure recovery factor, Cpj.

Cpi= Apj/0.5pV,2= 1 - o y v ,) 5

where Api is the ideal rise of static pressure across theblade row, and p is the air
density. To plot the curve for Cpj = 0.5 (or deH = 0.707)a numberofarbitrary
values of Oj are selected and the corresponding values of a, are calculated from:

a, = cos '(deRcosoij)

Knowing corresponding values of a, and a^ the flow coefficient can be found


from:

VaAJ = l/(tana, + tanap

The work coefficient follows from:

AH/U2 = 1 - 2tanOjVa/U

Table 3.1 illustrates the calculation for Cpj = 0.5. The same values of Oj as
used in table 3.1 may be used to make the calculations for other values of Cpj. A
series of these curves is plotted on figure 3.4.
The work andflow coefficient diagram 21

Table 3.1
Calculation of work and flow coefficients for a constant Cp,

“ 2° 0 15 30 45 60
45 46.92 52.24 60 69.3
VaAJ 1.0 0.748 0.535 0.366 0.228
AHAJ2 1.0 0.599 0.382 0.268 0.210

If a 50% reaction vector diagram is drawn corresponding to a point on one of


these lines and the blade speed is varied while keeping all other vectors of
constant length, as in figure 3.5, these vector diagrams correspond to a series of
points along a line of constant de Haller number. This must obviously be so
since deH = V^/Vj and V2 and Vj have been held constant.

Va/U

Figure 3.4 Lines of Constant Cpi on 50% reaction diagram

Vo- V2« Vo'- V2' V, - V ,- V ,'- V,'

Figure 3JS 50% reaction diagrams with equal velocities


but different blade speeds
22 Axialflow fans and compressors

What may not be so obvious is that the actual work is also constant That this is
so can be shown as follows:

« , - t 1- ( V ,, - V 22V2CF

where tj and ^ are the static temperatures at blade inlet and outlet respectively,
and Cp is the specific heat at constant pressure. Hence ^ ^ = At, will be
constant for all the diagrams with constant vector lengths but variable blade
speed. All the diagrams are for 50% reaction and:

R = At/AT; and hence: AT = At/R = a constant

It is a common practice to design for constant work at all radial stations along a
blade, so these diagrams could represent various radial stations of a stage. This
will be considered further when vortex flow is discussed.

Reactions other than 50%

To this point the discussion has been limited to 50% reaction. The AHJ\fi v
Va/U diagram can be drawn for other reactions or for vector diagram
arrangements such as 0^= 0°. A method of adapting the 50% diagram to other
reactions is to consider the rotor and stator separately as if each were part of a
50% reaction stage. Considering the vector diagram of figure 3.6, which is
clearly not 50% reaction, we can write:

AH/U2« 1 - (tanctjj + tana^Va/U

V>US30
Actual stage diagram: — ■■■ Equivalent 50% reaction diagram:
Urso = 50% reaction blade speed for equivalent rotor diagram.
Usso = 50% reaction blade speed for equivalent stator diagram.

Figure 3.6 Equivalent 50% reaction diagrams


The work andflow coefficient diagram 23

This gives a point on the AH/U3 v Va/U map but the values of Cpj and
deflection, plus other parameters yet to be introduced, cannot be read off at this
point. By considering the rotor alone as part of a 50% reaction stage a j and
would remain the same but the blade speed would have to be extended to the
right as indicated in figure 3.6. Dearly the values of AH/U and Va/U will be
less than for the true stage vector diagram, but the value of the parameter
AH/UVa = tan - tano^ will remain the same. Thus the equivalent 50%
reaction point on the work v flow coefficient map for the rotor can be found by
moving from the original stage point towards the origin along a straight line, i.e.
constant AH/UVa. The point required on this line can be determined by its
intersection with the line of constant representing the true rotor o^.
Alternatively, remembering that for 50% reaction Va/U = l/(2tanaai) the
equivalent 50% reaction value of the flow coefficient can be found and this
determines the appropriate point on the line of constant AH/UVa. The method is
illustrated in figure 3.7.
For the stator we proceed in a similar manner. In this case the equivalent blade
speed will be less than the true blade speed and the equivalent 50% reaction
point is found by moving away from the origin along the same constant AH/UVa
line as for the rotor. It will be apparent that lines of constant Cpj and AH/UVa
are tangential at Va/U = 0.5. If a non 50% reaction stage has its design point in
this region of the map, then the true values of Cpj will be less than indicated by
this point when the correction described above is made for the true reaction, and
this is the case for both rotor and stator.

VaAJ

Figure 3.7 Equivalent 50% reaction points for rotor and stator
24 Axialflow fans and compressors

The IGV/rotor fan

Taking Cpi = 0.5 as an approximate design limit it is instructive to compare the


curves for three different styles of fan design on the AH/U2 v Va/U map. On
figure 3.8 are shown a 50% reaction case, a zero design, and a zero a , design.
The 50% reaction case is shown only as a basis of comparison as this form of
design is seldom used for industrial fans due to it requiring both inlet and outlet
guide vanes, whereas the zero type requires no inlet guide vane and the zero
Oj type requires no outlet guide vane. It is clear from the curves of figure 3.8 that
in the lower range of flow coefficient, i.e. below Va/U = 0.5, the zero 013 type of
design permits a significantly higher level of work coefficient than either of the
other styles. This is its main advantage and its use tends to be confined to
designs requiring these relatively high levels of work. This ability of the zero
type of design is due to the high relative velocity at rotor inlet created by the
contra-whirl produced by the high deflection inlet guide vane. It is not suitable
for high performance aero engine fans or compressors because the high relative
inlet velocity would correspond to an unacceptably high Mach number.

VaAJ

Figure 3.8 Lines for Cpi = 0.5 for three types of fan

The rotor/stator fan

The zero 0^ type of design is of considerable interest as it is widely used for


single stage fans. Because of its wide application it is worthwhile constructing
the work v flow coefficient diagram for this case as in figure 3.9. For the outlet
stator the straight lines radiating from the origin are equal to constant values of
and therefore of the stator deflection for axial outlet conditions. The line for
tan otj = 1 is therefore a practical limit to stator deflection. At the other extreme.
The work and flow coefficient diagram 25

below OLy = 15° it is possible to eliminate the stator as the loss of the whirl
component of the outlet dynamic pressure is small. For the rotor the line of Cpj.
= 0.5 represents a reasonable upper limit to the area of practical design where it
is lower than the line of limiting stator deflection of 45°.
With uniform inlet total pressure the axial velocity is radially constant for this
type of design, at least in the absence of meridional curvature of the hub and
casing. The value of AH/Va is therefore of interest as this also is radially
constant when the work is radially constant, as is commonly the design
assumption. A number of such lines are shown and it will be noted that they are
much steeper than the constant Cp; lines. This means that the value of Cpj falls
quite rapidly from hub to tip which leads to a tendency for the optimum S/C to
rise rapidly and therefore the chord to diminish towards the tip of the rotor
blades. When the design point falls significantly below Cpj = 0.3 the S/C would
preferably be in excess of 2.5 and it may then be desirable to design on the basis
of isolated aerofoil theory rather than conventional cascade or compressor data.
For constant work AH/U = constant/r and it can be seen on figure 3.9 that if
the hub section is limited to Cpj = 0.5 and the tip section to not less than Cpj =
0.3 the diameter ratio will fall from 0.707 at AH/Va2 = 2 to 0.632 at AH/Va2 =
1.0. For AH/Va = 0.85 the curve only touches the Cpj = 0.5 line and within the
limits indicated the diameter ratio could be reduced indefinitely, or until the
o
limiting stator deflection of 45 is reached at the hub. However, in many
industrial fan designs it is advantageous to keep the axial velocity as low as
possible in order to avoid the need for downstream diffusion and its associated
losses and length requirements. This leads to a tendency to design at high values
of AH/Va in the region of 2. A low diameter ratio is also desirable in order to
minimise the outer diameter and hence low values of Cpj at the tip may be
inevitable. The line of Cpi = 0.5 intersects the limiting stator deflection at a
2
value of AH/Va = 3 and this represents a practical design limit
For the transonic fans used in high by-pass ratio aero-engines it is quite
common to design for a falling work and pressure ratio towards the hub. This is
because the low diameter ratios used would severely limit the by-pass pressure
ratio if the limited pressure ratio practical at the hub were maintained constant
along the whole blade span. Because the flow is split immediately behind the fan
with the core compressor inducing flow through the hub sections of the fan the
problems of a lower total pressure leaving the fan hub are largely avoided. As a
result the fan design pressure ratio is commonly of the order of 1.5 at the hub
rising to 1.8 or even 2.0 at the tip.
26 Axialflow fans and compressors

VaAJ
Figure 3.9 Work coefficient v flow coefficient diagram for oto = 0°
4 Blade loading parameters

Introduction

In order to achieve a satisfactory design it is necessary to select suitable vector


geometry which will permit the use of blade geometry of high efficiency and a
sufficient range of incidence between the design point and stall to allow for the
best achievable operating range. The design point must also allow a sufficient
margin from choke, which is dependent on the inlet angle and the incident Mach
number in relation to the throat width of the blade geometry. The determination
of parameters which control suitable vector and blade geometry has been the
subject of much theoretical and experimental research. Cascades of blades tested
in specially designed wind tunnels has been one of the principal experimental
techniques, since it offers one of the simplest and cheapest means of exploring a
wide range of blade geometries. While the data derived from cascade testing has
been of the utmost value to the compressor designer, it has required corrections
to be developed from tests of rotating machines. In that the cascade wind tunnel
can only be an approximation to the complex flow regime and blade row
interactions of a multi-stage compressor this is not surprising, indeed the surprise
is that cascade data has proved so very useful. It is only in recent years that
advanced computational methods have begun to replace reliance on cascade data
backed by analysis of research compressors.

Cascade data

Some of the earliest cascade data was produced by Constant, Howell, Carter, and
others working in British government research laboratories around the time of
the Second World War. Most of this data was presented in terms of ‘nominal*
conditions, defined as 80% of the maximum deflection achieved by a particular
cascade. The design data was presented as curves of deflection (e) against air
28 Axialflow fans and compressors

outlet angle (02) for constant values of the space/chord ratio (S/C). Other
presentations replaced deflection with Cpi or the theoretical lift coefficient, Cl*:
Ctih = 2(S/CXtanai - tana2)cosam.

An analysis based on diffuser performance resulted in an analytical expression


giving almost identical results:

Clu>(V|/V2)3 = 2.2

Similar work carried out in the U.S.A. by staff of NACA, the predecessor of
NASA, resulted in a blade loading parameter known as the diffusion factor. This
was derived from consideration of the diffusion taking place on the convex
(suction) surface of the blade from the point of maximum velocity to the trailing
edge. The diffusion factor is given by:

Df = (1 - V2/VO + AVw(S/C)/2V,.

For incompressible two-dimensional flow this can be written:

Dp = (1 - cosai/cosa2) + 0.5(S/C)cosai(tana, - tana2)

Va/U
Figure 4.1 Howell's nominal deflections in terms of 50% reaction
work and flow coefficients

Values of loss parameters were correlated against the diffusion factor and
indicated that the loss rose rapidly above a value of Df = 0.6. Howell's nominal
deflection data can readily be shown to be approximately equivalent to a
diffusion factor of 0.45 and his nominal design rules are shown in terms of the
Blade loading parameters 29

work and flow coefficient diagram on figure 4.1. The region between the curves
for S/C = 0.5 and 1.5 effectively defines the region of practical design.

Total pressure losses

Only two of the many analyses of cascade pressure losses will be described;
those due to Howell (194S) and Lieblein (1965).
Howell presented data for losses in the form of the profile drag coefficient*

Cdp = GJiS/C cos3am/cos2a,

where a* is the vector mean angle, <X| is the inlet air direction, ffl, is the total
pressure loss coefficient AP/VipV,2 and S/C is the space chord ratio. The
minimum value of Cop for all cascades is given as 0.018 for Reynolds numbers
above the critical value of 3*10*.
For the total losses, values of annulus friction drag, Cd. and secondary drag
Cd«are added to the profile drag:

Cd = Cop + Cd» + Cd,

where Cd. = 0.02S/h and Cd, = 0.018Q.2, h being the blade span and Cl is
usually taken as the theoretical lift coefficient as previously defined. The total
pressure loss coefficient is derived using Cd in place of Cop in the previous
relationship.
Lieblein correlated the losses of low speed cascade losses in terms of the
diffusion factor Dp. He presented graphs for two loss parameters. The simpler
one gave a linear relationship with diffusion factor which can be written as:

05, = 2(Df + 0 .3)/100(S/C)cosa2

This is equivalent to Howell’s profile loss and is stated to apply for Reynolds
numbers between 2*103 and 2.5* 105. Other losses have to be derived from
another source such as data to be discussed in Chapter 6 from Wright and Miller
(1991) or Howell's annulus and secondary drag parameters:

Cd. = 0.02(s/h) and Cd, = 0.018CL2

Deviation

Blade design depends critically on an accurate knowledge of the air outlet angle
which a given blade geometry will produce at the nominal or minimum loss
incidence. The angle difference between the air outlet direction and the blade
30 Axialflow fans and compressors

outlet angle is known as the deviation, 8 . The blade outlet angle is the angle
between the tangent to the blade centre line and the axial direction (see figure
1.2b). The British data was correlated by a rule known as Carter's rule, which is
of the form:

8 = meV(S/C)

where m is a factor dependent on the blade stagger angle, 6 is the blade camber,
equal to the difference between the blade inlet and outlet angles. The deviation
values given by American data are generally greater than those given by Carter's
rule by amounts up to 3°. A generally accepted explanation for this is that the
American cascade tests were conducted with porous walls at the blade ends to
which suction was applied to prevent growth of the wall boundary layers. The
early British data was obtained without wall suction, with the result that the wall
boundary layer growth caused an acceleration of the axial velocity through the
midspan section of the cascade. This reduces the static pressure rise across the
blade and increases the air deflection, thus reducing the deviation. The outcome
was that a work factor was necessary when designing with the British data. This
was such that the vector diagram work was increased by about 15% over the
actual work required. At the time it was claimed that the work factor was
necessary to account for the development of wall boundary layers through the
stages, leading to higher than intended axial velocities at mid-annulus causing
reduced work. In the American system a blockage factor is introduced to account
for this effect and most designers have now adopted this method.

Axial velocity ratio

The change of deviation angle with variation of the axial velocity ratio across the
cascade can be a significant factor and this is commonly included in the design
process. An approximate rule given by Gostelow (1984) is:

A5 = k(l - Vaj/Va,)

where k lies between 5 and 10, the value tending to the larger number at high
values of the stagger angle. Thus for an axial velocity ratio of 0.8 the deviation
may be as much as 2° greater than for a velocity ratio of unity, but it is probably
satisfactory to assume k = 7.5 for most cases.

Compressor data

A series of tests conducted at Rolls Royce on a small low speed research


compressor gave support to cascade design data. This work was fully described
Blade loading parameters 31

in McKenzie (1980) and McKenzie (1988); only a summary is given here. A


notable result, for the 50% reaction stages used in the experiments, was that the
blade stagger angle appeared to deteimine the flow coefficient for maximum
stage efficiency, independent of other blade geometry features. This can be
interpreted as a relationship between stagger angle, £, and vector mean flow
angle, On, and is then independent of reaction.

tanOn, = tan£ + 0.213

5°AVR
Figure 4.2 Compressor and cascade deviation
From McKenzie (19SS) with permission of I Mech E

At low staggers the maximum efficiency occurs close to stall and to provide a
stall margin an alternative design rule is:-

tanctn, = tan£ + 0.15.

Although cascade data had not been correlated in this manner it was found that
data obtained at Rolls Royce from a series of low Mach number cascade tests
with wall boundary layer suction agreed well with the above relationships. Two
deviation rules were derived from the compressor data. The first was derived
from direct measurement of the airflow direction leaving the first stage stator.
This gave:

5,=(x.i + 0.31exs/C),/3
The second rule was derived from the overall work input and mass flow of the
compressor and gave the relationship:

5w = (2 + e/3)(S/C)1'3

This gives values approximately l°to 2.5° greater than 8,, and this difference is a
measure of the blockage factor of this compressor. The 5, values agreed within
32 Axialflow fans and compressors

about ± 1° with the same cascade data, 5Avr» both giving values about 2° greater
on average than Carter's rule, and therefore very similar to the NACA deviations.
These deviation rules are different from some others in that they are independent
of the stagger angle, and have been criticised for this reason, as theoretically
deviation should vary with stagger. Stagger angles of 20°, 30° and
50° were used in the experiments and figure 4.2f taken from McKenzie (1988),
shows the modest experimental scatter.

Blade geometry selection

A correlation for design selection of efficient blade geometry was produced from
the same experimental compressor in the form shown in figure 4.3, taken from
McKenzie (1988). Contours of constant efficiency are plotted on a field of Cpi v
S/C. As might be expected the optimum efficiency line has a falling value of Cpi
as S/C increases. Efficiency falls off if Cpi is greater or less than the optimum
value for a given S/C. The efficiencies quoted on figure 4.3 are the values
obtained from the small scale compressor tests, and are therefore appropriate to

i/c

Figure 43 Efficiency contours


From McKenzie (1988) with permission of I Mech E

the test conditions of Reynold's number, Mach number, tip clearances, etc. They
are therefore not directly applicable to other situations, but provide a guide to the
relative efficiency potential of the chosen blade geometry. The same data is
presented in the form of the loss coefficient in figure 4.4.
Blade loading parameters 33

«c
Figure 4A Contours of loss coefficient derived from figure 4.4

s*c

Figure 4>5 Range of blade geometries for Oi = 60° and a 2 = 50°


From McKenzie (1988) with permission of I Mech E

Since it is not always practical to choose the optimum blade geometry this
method allows an informed choice as it indicates the loss of possible maximum
efficiency involved by the chosen geometry. The following example illustrates
the selection of possible blade designs for ai = 60°, and a 2 = 50°.

tan On = 0.5( tan 60° + tan 5(f) = 1.4619

tan ; = 1.4619 - 0.15 = 1.3119 = 52.7°

Assume i = a r Pi = - 5°, where pi is the blade inlet angle.

Pi = 60° + 5° = 65°.

0 = 2(p, - 0 = 2(65° - 52.7°) = 24.6°


34 Axialflow fans and compressors

P2 = p, - 6 = 65° - 24.6° = 40.4°

8 = a ,- P, = 50° - 40.4° = 9.6°

= (1.1 +0.31*24.6XS/C)W

Hence: S/C = 1.33

Cpi = 1 - (cos2ai/cos2ct2) = 0.395

From figure 4.3 r\ = 90%

Repeating the above calculation for other values of incidence in the range from
-10° to +5° gives the results plotted in figure 4.5.
5 Blade geometry

Introduction

In the UK until recent years compressor blades were most commonly based on a
specified uncambered aerofoil to define thickness at a series of chordal stations.
The thicknesses were scaled to provide any desired ratio of maximum thickness
to chord. Usually this would lie between 5% and 15%. These thicknesses were
then set off on a chosen camber line, which was either a circular arc or parabolic
arc. The parabolic arc generally had its maximum offset from the chord line at
40% of the chord from the leading edge. Although some designers preferred the
parabolic camber line, experience suggested that there was no major advantage
of one over the other. Because of its simpler geometry the circular arc camber
became the more popular.

The C series aerofoils

Basic geometry features for a C4 aerofoil are illustrated in figure 5.1. The C5
profile is identical to the C4 from the leading edge to 40% of the chord and then
tapers to a trailing edge radius of 12% of the maximum thickness compared to
6 % for the C4. The C5 is used on small chord blades where the thinner trailing
edge of the C4 is impractical. There is little additional loss with the C5, and C4
design data is satisfactory for both profiles.
The C7 profile was intended for higher subsonic Mach number applications
and has the maximum thickness moved back from 30% to 40% chord. It has not
been used extensively because the Double Circular Arc (DCA) blade section,
described below, has been found superior for high Mach number applications.
36 Axial flow fans and compressors

-a —■
— '—
0 10 20 30 40 50 60 70 80 90 100
Chord %

%Chord 1.25 2.5 5.0 7.5 10 15 20 30


0.5t/C% 1.65 2.27 3.08 3.62 4.02 4.55 4.83

%Chord 40 50 60 70 80 90 95
0.5t/C% 4.89 4.57 4.05 3.37 2.54 1.60 1.06

Leading edge radius = 12% of maximum thickness.


Trailing edge radius = 6% of maximum thickness.

Figure 5.1 C4 aerofoil base profile


Adaptedfrom Howell (1945) with permission of I Mech E

American practice

In the U.S.A. the NACA 65 series of aerofoils was most commonly used. These
approximate to a circular arc camber line but have a somewhat different
thickness distribution to any of the British C series. The maximum thickness of
the NACA-65 series is at 40% chord and the trailing edge thickness is
theoretically zero. For obvious practical reasons this was commonly modified to
a radius of 0.8% of chord, which gives the same trailing edge thickness as a C4
or C7 profile when the maximum thickness to chord is 13.3%, and the same as a
C5 at 6.7%.
Details of the C series of aerofoils are given in Howell (1944) and an appendix
to Cumpsty (1989) gives details of both the C series and the NACA 65 aerofoils.

Cascade geometry

Apart from the aerofoil geometry itself there are two important parameters which
relate the blade to its neighbours, and define its orientation in the compressor.
Blade geometry 37

These are the ratio of the circumferential spacing to the chord, S/C, and the angle
of the chord line to the axial direction, known as the stagger angle, £.

Figure Cascade geometry

In American usage C/S is commonly used rather than S/C and is referred to as
the solidity, generally denoted as a. Industrial fan blading usually defines the
blade angles as 90° - £, i.e. the angle is defined from the peripheral direction
rather than the axial.

The double circular arc blade

The C series and, to a slightly lesser extent, the NACA 65 series tended to show
a rapid rise of pressure loss at Mach numbers above 0.7 to 0.8 depending on the
t/c, camber and S/C. A very simple blade form which proved to be superior at
high subsonic and transonic Mach numbers is one in which both surfaces are
formed by circular arcs. The leading and trailing edges are formed by radii equal
to 6% of the maximum thickness. This is known as a Double Circular Arc
(DCA) blade. The camber line is obviously also a circular arc, and the maximum
thickness is at 50% of the chord. A comparison of the cascade performance of
DCA and C4 blades is given in Andrews (1949). At Mach numbers below the
critical there is no significant difference between the performance of a DCA and
a C4 blade of the same geometry other than the base profile. This is
demonstrated by figure 5.3 where the results of low speed tests on two sets of
blades, identical except for the base profile, show negligible performance
difference. The same design rules can be used to choose S/C, camber and stagger
38 Axialflow fans and compressors

for DCA and C4 profiles. DCA profiles can be used succesfully in the transonic
range for the design of rotor blades up to a Mach number of 1.2.
Of the blade profiles discussed thus far there would seem to be no performance
reason to use other than the DCA blade profiles. The more conventional aerofoil
sections may have some advantage in their robust leading edge shape in terms of
erosion and foreign object damage provided the Mach number is not excessive.

v*u

Figure 53 Performance of C5 and DCA blades at low Mach No.

Modern trends

The introduction of the supercritical aerofoil for aircraft wings and the rapid
growth in computing power led to the introduction of compressor aerofoils
designed for a prescribed surface velocity distribution from the late 1970’s. This
design method produces an effectively tailor made blade profile for each design
requirement. If a cascade of aerofoils is to be designed for a predetermined
velocity (or pressure) distribution it is essential that this should be defined in
such a way as to achieve the minimum loss practical. The primary requirements
are to have a smooth and continuous acceleration from the leading edge to the
point of maximum Mach number on the suction surface, followed by the
maximum rate of diffusion possible without flow separation. It is also necessary
to avoid a shock wave arising, and this limits the maximum suction surface Mach
number to no more than 1.3. These desirable characteristics are illustrated in
figure 5.4. It is not usually practical to define the desired velocity distribution
and calculate the required blade shape directly. A first approximation to the
aerofoil has to be defined, its velocity distribution is then calculated and the
profile modified to give a closer approximation to the desired velocities.
Blade geometry 39

Smooth
acceleration

% Chord

Figure 5.4 Mach No. distribution round controlled diffusion blade


Adaptedfrom Hobbs and Weingold (1983) with permission of ASME

Obviously this process can be time consuming and experience is necessary to


iterate rapidly to the final blade profile. An inviscid flow is assumed in most
forms of the calculation and the initial shape defined includes the surface
boundary layer displacement thicknesses. These are then calculated, for example
by the methods of Stow (1984) and subtracted from the inviscid blade
thicknesses to give the actual blade form.

Figure 5.5 Comparison of typical equivalent aerofoils

The aerofoil shapes which result are distinctly different to more conventional
shapes, as shown by the comparison between a DCA and controlled diffusion
profile for the same duty in figure 5.5. The principal features of the latter are the
relatively large leading edge radius and the camber line shape which has
increasing curvature from the leading edge to 30% or 40% of chord, after which
the curvature diminishes to give almost no camber in the latter 20% to 30% of
the chord. The straightness of this part of the blade plus the lack of separation
40 Axialflow fans and compressors

from the suction surface account for the low deviation of the exit air angle from
the blade trailing edge angle. Typically this will only be 2 or 3 degrees compared
to 5 to 10 degrees for conventional blade profiles. Predictions of the deviation
have proved at least as accurate as cascade data is for conventional blades, and
successful designs can be achieved without resort to experimental data, of which
relatively little exists in the open literature.
Stow (1984) shows that the difference in suction surface shape to remove a
shock and achieve shock free diffusion on the suction surface can be quite small,
e.g. a difference in thickness of 0.4mm at most for a blade of 50mm chord and
3.2mm maximum thickness. Manufacturing accuracy must therefore be of a high
order, and any erosion of the blade form in service may lead to a serious
deterioration of performance.
The maximum Mach number for which the controlled diffusion profiles are
suited are not significantly different to those of C4 or NACA 65 blading. Since
the high subsonic Mach number performance is largely determined by the blade
shape from the leading edge to the throat and the throat width itself, this is not
surprising since the controlled diffusion profiles are more similar there than
towards the trailing edge.
The work involved in defining controlled diffusion profiles for the large
number of sections required for a multistage compressor is very considerable.
This has led to attempts to define families of shapes, as for the conventional
blade definitions. This has not been successful except in so far as providing a
useful first approximation for the iteration of the final profile definition. The
performance gains from the introduction of this type of blade profile are
confined to relatively small areas and have thus far been perhaps less than was
anticipated from early indications. None the less they have become popular with
design organisations having sufficient computing power to carry out the large
amount of calculation involved.
Tubbs and Rae (1991) indicate that controlled diffusion blades can be prone to
vibration due to the low trailing edge wedge angle of this type of blade. In the
development of the high pressure compressor for the IAE V2500 aero engine
they state that a compromise profile between the controlled diffusion aerofoil
and conventional circular arc sections proved the best compromise on
performance and integrity.
6 Aspect ratio

Introduction

Blade aspect ratio is defined as the ratio of blade height to chord. The height is
measured at mid chord, or can be taken as the mean of the blade span at leading
and trailing edges. The chord is commonly defined as the value at mid span.
Typical values for modem compressors lie between 1 and 2.5 but were often
higher in designs prior to 1960 before the benefits of lower aspect ratios were
fully appreciated. The history of the subject in America has been described by
Wennerstrom (1986). In Britain the early work by A.R.Howell suggested that
higher aspect ratios would only reduce the annulus drag coefficient, and this
indicated a marginally higher efficiency for higher aspect ratios. As a result
aspect ratio was not a major aerodynamic consideration. Its value was generally
decided by mechanical limits of stress and vibration, with minimum overall
length for a given number of stages as the primary objective. The result was
many designs which barely met their design pressure ratios, or had inadequate
surge margin, as described by Wennerstrom (1986) for American practice.

Surge margin improvement

The origin of awareness of the benefits of lower aspect ratio blading at Rolls
Royce has been described in McKenzie (1984) and is repeated here, because it
illustrates dramatically the influence of aspect ratio on surge pressure ratio from
otherwise similar blading.
When design commenced for the Rolls Royce Tyne turboprop engine, it was
decided to base the HP compressor design on that of the Conway HP, which had
been developed to a satisfactory performance standard. This required the
Conway blading to be scaled down by a linear factor of 0.54. When this was
considered the blade chords were judged to be too small, both on account of
manufacturing accuracy, and Reynolds number. As a result the chords were
42 Axialflow fans and compressors

significantly increased, to the extent that the final stage rotor aspect ratio was
reduced, in approximate numbers, from 2 to 1, and the overall length was
increased by about 50% over that of a true scale of the Conway HP compressor.
The number of blades in each row was reduced as necessary to keep the same
values of S/C as the Conway. Comparative annulus diagrams are shown in figure
6.1, and the overall performances are compared in figure 6.2. The additional
surge margin provided could only be attributed to the lower aspect ratio blading.
Naturally this led to extensive research into the effects of aspect ratio, and from
the mid 1960's lower aspect ratio blading was more generally adopted.

" '' 1


R S F OG\
1 1 9 t

0.54 linear scale of Conway HP compressor

Figure 6.1 Tyne HP reduced aspect ratios

Other benefits

The use of lower aspect ratio, together with the more widespread use of titanium
alloys for blading, had advantages beyond the immediate one of increased surge
pressure. The combination of improved material and lower aspect ratio allowed
the use of higher blade speeds. In consequence an increase of flow Mach number
was possible without loss of efficiency because of the lower thickness to chord
ratios which were permissible. The adoption of double circular arc aerofoils
completed a dramatic improvement in axial compressor engineering. The total
number of blades required to produce a given pressure ratio has been reduced by
a factor in the region of 4, while efficiency has been maintained and, where
desired, mass flow per unit area has been increased. Hill, Nicholas, and Tubbs,
(1990), give further historical details of these developments.
Aspect ratio 43

%Dosign mass flow

Figure 62 Conway HP and Tyne HP performance comparison

Figure 6.3 Reduced aspect ratio stage

Technical considerations

In aircraft wing performance it is well established that higher aspect ratio


produces improved efficiency as measured by the ratio lift/drag. Not surprisingly
therefore it was expected to have a similarly beneficial effect in axial
44 Axialflow fans and compressors

compressors. In ternis of stage maximum efficiency the evidence is not clear.


Some comparative tests indicate a small gain at higher aspect ratio, but just as
many indicate a loss. A rule of thumb is to assume maximum efficiency does not

Va/u

Figure 6.4 Performance effect of tip clearance and blade height

change with aspect ratio. The increase of stall or surge pressure ratio at lower
aspect ratio is thoroughly established, and is a significant part of the designer's
ability to achieve a given pressure ratio in no more than half the number of
stages required thirty years ago.
L.H.Smith (1969) produced the first satisfactory analysis of aspect ratio
effects. He argued that the influence of annulus wall drag, tip clearance and
secondary flows would give an end wall boundary layer displacement thickness
which could be related to the staggered gap (Scos£ ) and the tip clearance. The
boundary layer is expressed in terms of the axial velocity profile as it falls off
towards the end walls. While it may be argued that the true boundary layer
should be defined in terms of total pressure or the total velocity component
relative to the wall concerned, this in no way alters Smith's analysis.
The principle of his argument is explained with the aid of figure 6.3. The
repeating stage assumptions, as described in Chapter 13 are invoked. The radial
distribution of axial velocity repeats after each stage, and the static pressure rise
is equal at all points along the blade height at a given operating point. If now a
part of the blading near the centre of the span, where the flow is uninfluenced by
the wall effects, is removed, then the stage aspect ratio is reduced, but the size of
the secondary flows will remain unchanged. In order to produce the same
pressure rise the axial velocity must be the same as for the original stage at mid­
span where the flow is two dimensional and the total losses are the blade profile
Aspect ratio 45

losses only. Due to the unchanged thickness of the axial velocity boundary layer
this pressure rise will occur at a lower mass flow or average flow coefficient.
The stall flow will be similarly reduced, and the pressure rise will be lower,
despite the lower aspect ratio, because of the greater proportion of the blade
height occupied by the end losses. If the tip clearances are reduced to the original
proportion of the blade height the efficiency and pressure rise will increase. As
indicated in figure 6.4 the stage characteristic will now show the typical effect of
a reduced aspect ratio compared to the original characteristic.

Equivalent cone angle

Schweitzer and Garberoglio (1984) draw a comparison of the compressor


cascade to a conical diffuser. They correlated cascade data with a parameter
defined by the ratio of the maximum achieved static pressure rise coefficient to
the ideal coefficient at the maximum efficiency operating condition. Although
they called this parameter the ‘maximum pressure rise efficiency', it is perhaps
more appropriately considered as a stall margin parameter. Experimental values
of the parameter plotted against the equivalent conical diffuser angle indicate a
maximum stall margin at an equivalent diffuser angle of approximately 3.5°,
which agrees well with established diffuser data. The equivalent conical diffuser
angle can be defined with reference to figure 6.5 as:

tan a ,, = {(h/C)(S/CXl/K)),n{cos,<V c o s 'na l )

If Howell's curves of nominal deflection are used to give related values of ai,
a 2 and S/C, it is found that, for an assumed aspect ratio of unity, they
consistently give values between 3° and 4° for the equivalent cone angle. This
would suggest that if maximum stall margin is desired for a design based on
Howell's data, and an aspect ratio greater than unity is required, the S/C should
be reduced below that indicated by Howell's data until an equivalent cone angle
of 3.5° is obtained. This will occur when (h/CXS/C) = (S/C)h, where (S/C)h is
Howell's recommended S/C. No direct experimental evidence can be offered for
this suggestion, but it is the case that many higher aspect ratio early designs
based on Howell's data were deficient in surge margin.

End wall loss parameter

Freeman (1985) gives a curve of diffusion factor near stall as a function of the
ratio of dp clearance to chord. More recently Wright and Miller (1991) include
the influence of aspect ratio in an ‘end wall loss parameter' This parameter is
plotted as a series of curves for constant values of the ratio clearance to chord
against the diffusion factor, as in figure 6.6. The loss parameter rises rapidly for
46 Axialflow fans and compressors

each curve above a critical value of the diffusion factor which increases as the
clearance to chord reduces. Freeman's curve would appear to be related to this
data at a value of approximately 0.1 for the end wall loss parameter as indicated
on figure 6.6. A possible design area would be the region between the two curves

R-(A/it)'n ; r = (a/n)'n

a a s.h.cosai
A = s.h.cosct:
L **C
tanas {(l/jcXh/CXS/QJ^cos^ai-cos,/2a,)
Figure 6.5 Equivalent conical diffuser angle

indicating 80% and 90% of the stall diffusion factor. For e/C between 0.02 and
0.04 this again indicates the design diffusion factor should be close to 0.45.
Aspect ratio 47

If the ratio of clearance to blade height is assumed constant at, say, 2%, then
the ‘endwall loss parameter’ curves can be considered curves of constant aspect
ratio because: h/C = (e/C)/(e/h). The values of e/C plotted give values of aspect
ratio as shown in table 6.1
If the minimum e/h is determined from mechanical considerations, including
such influences as transient operation and manufacturing tolerances, then an
appropriate choice of diffusion factor and aspect ratio can be made.

Table 6.1
Clearance to chord and aspect ratio

e/C 0.1 0.07 0.04 0.02


h/C 5 3.5 2 1 for e/h = 0.02
h/C 2.5 1.75 1 0.5 for e/h = 0.04

Diffusion factor

Figure 6.6 Wright and Miller’s end wall loss parameter


From Wright and Miller (1991) with permission of I Mech E

Design criteria

It may be asked what design recommendations are to be assumed regarding the


choice of aspect ratio. There is no direct connection between the various sources
of data referred to, and the designer must either make a choice between them or
attempt to achieve a design which satisfies more than one set of data. For
example a design diffusion factor may be selected in line with the minimum
clearance judged practical, the required blade height, and a suitable aspect ratio.
The effective cone angle can then be checked against Schweitzer and
Garberoglio's data for adequate stall margin. There is a tendency in most
48 Axial flow fans and compressors

compressors for the aspect ratios to reduce towards the rear stages. Greater stall
margins are therefore available on the rear stages than on front stages. This
provides greater surge pressure ratio at high speeds, where rear stages approach
stall first. To provide improved surge margin at low speeds would require low
aspect ratio for the front stages, but since the blade height is greater for these
stages, this would lead to very long chords and increased overall length.
Furthermore, adopting the argument that clearance to chord is the most
significant parameter, this can usually be kept to a low value on front stages
where chonJs are usually larger than for rear stages, temperature effects are
smaller, and diameters are of the same order as for rear stages. Where design
pressure ratio is in excess of 6 to 1, variable stators are often adopted on the
front stages and this reduces the requirement for large stall margins on the early
stages.
7 Relative motion effects

Introduction

The relative motion of succeeding blade rows has a most beneficial effect on
axial compressor performance. It may be no exaggeration to suggest that these
machines would be impractical if it were not for this phenomenon. The effect is
illustrated in figure 7.1, where a vector diagram is drawn in which <Xj + Oo = 90°
for the main stream flow. If the reduced velocity flow leaving the upstream stator
close to the casing is considered, it is apparent that while the velocity in the
stator frame of reference may be half of its mainstream value, the velocity
relative to the following rotor is little different to the mainstream value, although
its angle of incidence to the rotor has increased significantly. Figure 7.2 shows
velocity diagrams in which ai + cto is much greater and much less than 90°.

V „ = Main stream stator


exit velocity

VM s Stator exit boundary


layer velocity

Figure 7.1 Constant relative inlet velocity for cto + ai = 90c

In contrast, if the situation for mainstream and wall boundary layer flow into a
cascade is considered, because both flows have the same incident angle, there is
a large difference in the incident velocities of the mainstream and the boundary
layer. If the mainstream velocity is diffused in the blade passage to 0.8 of its
inlet value, then ignoring losses, it is found that for an inlet velocity of 60% of
50 Axialflow fans and compressors

the mainstream value the whole of the dynamic head would require to be
converted to static pressure to give the same static pressure rise as the
mainstream. The static pressure rise is therefore limited to 36% of the
mainstream inlet dynamic pressure, i.e. 1 - 0.82 = 0.36. This is typical of the
maximum static pressure rise achieved by cascades when no boundary layer
suction is applied to the walls. There is, in reality, a transfer of energy from the
mainstream to the boundary layer and the real flow is more complex than
assumed in the above explanation, however it does serve to illustrate the greater
limitation there would be to the pressure rise achievable by a compressor stage if
it were not for the relative motion effect

Figure 7.2 Velocity diagrams for Oo+ ai > and < 90°

It is clear from the performance achieved by compressors that the velocity ratio
across a blade row may be as low as 0.7, rather than the 0.8 limit of a cascade
operating without wall boundary suction. In the casing and hub boundary layers
this is achieved because of the relative motion effect, but at the expense of an
increase of incidence. It is well established that there are considerably higher
losses of total pressure close to the annulus walls than at blade midspan. It is
possible that the blades will accept a higher incidence in the presence of high
losses which are not induced by the incidence itself, i.e. such as those due to
wall friction and the secondary flows thereby induced. A hypothesis to support
this is given below.

Blade design with secondary losses

Howell (1945) presented design data in terms of the theoretical lift coefficient,
and alternatively the ideal static pressure recovery factor. This is satisfactory
when the lift/drag ratio is 50 or more, and the total pressure loss coefficient less
than about 0.05. When the losses are much larger it is unlikely that they can be
ignored. If the situation is considered in terms of the pressure coefficients:

CPi= l - ( V 2^ /i)2

Cpact ~ Cpi" G3
Relative motion effects 51

where OS is the total pressure loss coefficient Let an effective ideal pressure rise
factor be defined as:

Cpeff = Cpact + Op

where 0 P is a nominal loss coefficient due to blade profile losses only, and
assume for illustration purposes (3P = 0.025 and GJ = 0.125. For constant axial
velocity through the blade row:

Cps = 1 - (cos2ai/cos2a 2)

and Cpeff = 1 - (cos2aieff/cos2a 2)

This defines an effective inlet air angle, cti^r, and the blade geometry can be
selected to suit this angle, and the required oudet angle, using conventional
design data.
For example: Let ai = 58° and a 2 = 36°; 0 = 0.125 and 0 P= 0.025

Cpj = 1 - (cos258°/cos236°) = 0.571

Cp^ = 0.572 - 0.125 = 0.446

Cpdf = 0.446 + 0.025 = 0.471

Hence: 0.471 = 1 - (cos2a leff/cos2 36°)

and (Xieff = 54°

Therefore the blade design assumes an inlet angle of 54° instead of 58° and an
outlet angle of 36°. Following the design procedure described in Chapter 4:

tana,,, = 0.5(tan54° + tan36°) =1.051

tan; = 1.051 -0.15 = 0.901

C = 42.0°

Choose i = -5°.; hence pi = 59° and the camber, 0 = 2(59 - 42) = 34°

a 2 = 59 - 34 = 25° and the deviation, 5 = 36 - 25 = 11°

Deviation is also given by: 8 = (1.1 + 0.31* 0 )(S/C)1/3


52 Axialflow fans and compressors

Hencewehave: 11 =(1.1+0.31* 34XS/C)10

From which : S/C = 0.844

The actual incidence, i= 58 - 59 = -1°

Provided the actual incidence is within the normal range -5° < i < +5° the blade
operation should be satisfactory.
Applying the same design rules directly to the vector angles, a t = 58°; 02 =
36°, without allowance for the large secondary losses, results in the range of
possible blade geometries listed in Table 7.1.

Table 7.1
Blade camber and S/C for a range of incidence

1° 0° S/C
0 25.2 0.046
-5 35.2 0.318
-10 45.2 0.666

This indicates that, without allowance for high secondary losses, a very low
S/C must be used, or an undesirably large negative incidence is required.
Despite the lack of direct experimental evidence for this hypothesis, it can be
claimed that it recognises the real pressure forces on the blade to a greater degree
than the more traditional design methods. The assumption of a false incidence
angle may be thought inappropriate, but an alternative view is to consider the
operation with high secondary losses as being analogous to operation with an
accelerating axial velocity through the blade passage. For the same inlet angle,
the acceleration of the axial velocity reduces the pressure forces on the blade
similarly to the presence of large secondary losses. Experimental results for
cascades with variation of the axial velocity ratio across the blades indicates that
stall occurs at an increasing incidence as axial acceleration is increased. At a
constant incidence the deviation of the outlet angle also reduces as the axial
acceleration increases. These results are considered to add some credibility to the
hypothesis.

Wakes in relative motion

It is obvious that at any spanwise position the reduced velocity occurring in the
trailing wake of a blade will give rise to the same relative motion effect on the
inlet velocity to the succeeding blade row as a hub or casing boundary layer. In
this case the downstream blade is passing through the wakes of the upstream
Relative motion effects 53

blades at a frequency equal to the product of the number of blades in the


upstream row and the revolutions per second. The downstream blade is therefore
subjected to a rapid fluctuation of incidence, but a substantially constant
velocity. Dring and Spear (1990) have shown that the blade tends to respond to
the fluctuating flow as if it were a steady flow having the mixed out velocity and
angle which can be deduced from a momentum mixing calculation.

Work and flow coefficient effects

When the sum of the absolute and relative angles is 90° a small change in the
absolute velocity will produce no change in the relative velocity at rotor inlet At
stator inlet the same applies, but it is a small change in the velocity relative to the
rotor outlet that is considered. As shown by figure 7.3 for 50% reaction these
conditions occur at VaAJ = 0.5 when AH/U2 = 0; and Va/U falls to 0.433 when
AH/U2 is 0.5. For larger values of Va/U the sum of the angles is less than 90° and
the inlet velocity to the following blade row is not fully restored to its
mainstream value. When Va/U is less than these values, and the sum of the
angles is greater than 90° the boundary layer or wake value of the inlet velocity
is greater than the mainstream velocity. There is a general tendency for
efficiencies to be higher and stall margins to be greater at low design flow
coefficients, and the relative motion effects may well contribute to this.
At reactions other than 50% the condition for the sum of the angles to be 90°
vary somewhat For example, for designs having zero whirl at stator exit the sum

VaAJ

Figure 73 Sum of angles greater and less than 90°


54 Axialflow fans and compressors

of at and Oo is always less than 90°. As also shown on figure 7.3, however, at
stator inlet of such designs the sum of the angles is 90° at VaAJ = 0.5 and AHAJ2
= 0.5, and the line for a 2 + a 3 = 90° lies in the conventional design area below
this point. Thus the conditions for the coefficients to give the sum of the angles
equal to 90° are not dissimilar to the 50% reaction case, and broadly there would
appear to be an advantage to designs having a flow coefficient of 0.5 or less.
TTiere are, of course, other compelling reasons why this value of the flow
coefficient is exceeded in certain cases, for example near the hub of inlet stages
of low diameter ratio.

Radial flow shift through the rotor

Another effect of the modon of the rotor is the work done on the flow when it
moves to a greater radius as it passes through a rotor, or conversely the work
extracted from the fluid if the radius decreases in passing through a rotor. If the
radius of a streamtube remains constant through the rotor the total temperature of
the flow relative to the rotor is constant If the stream tube radius increases the
relative total temperature is increased. This is most commonly of concern for
flow near the hub where the radius increases towards the rotor trailing edge.
Conversely, if the flow migrates to a lower radius as it passes through the rotor,
as for example along a tapered casing, the relative total temperature decreases.
Of course, there are no similar effects in stators due to the lack of any rotational
motion.
The increase of enthalpy across the rotor is:

AH = H3 - Ho = U2Vw3 - U, Vw0

where Ui and U2 are the blade speeds at rotor inlet and outlet respectively and
Vw0 and Vw3 are the absolute whirl velocities at inlet and outlet respectively.
The enthalpy relative to the rotor inlet is:

H, = Ho + Ui(Vw, - Vw0)/2

where Vwi is the whirl velocity relative to the rotor at inlet.


Relative to the rotor at outlet the enthalpy is:

H2 = H3 - U2(Vw3 - Vw2)/2

where Vw2 is the whirl velocity relative to the rotor at outlet.


From which the difference in total enthalpy relative to the rotor from inlet to
outlet can be derived as:

AHrej = H2 - H, = (U22 - U,2)/2


Relative motion effects 55

This increase of enthalpy would be present even if there were no fluid velocity
relative to the rotor, therefore the increase of relative total enthalpy is not subject
to flow losses and the accompanying increase of pressure is effectively achieved
at 100% efficiency.
8 Vortex flow

Introduction

The distribution of tangential, or whirl, velocity with radius is an important


design consideration. If the meridional streamlines form concentric cylinders the
flow is said to be in simple radial equilibrium and the radial variation of axial
velocity will depend on the radial distribution of tangential velocity.

Free vortex designs

A free vortex is the type of flow pattern which arises naturally such as water
draining from a bath tub. It is characterised by having constant angular
momentum at all radii, so that:

Vw.r = a constant

We also have the relation developed in Chapter 3:

Vw0 = (l-R)U - AH/2U

and since U = cur, after some manipulation we obtain:

R = 1 -k/r2

Putting r = 1 where R = 50% the variation of reaction with radius is given in


Table 8.1 The figures show that the reaction approaches unity asymptotically as
the radius increases. Note that reaction rises from 50% to 87.5% when the radius
is doubled.
Vortexflow 57

Table 8.1
Variation of free vortex reaction

r 0.707 1.0 1.414 2.0 2.5 3.0


R 0 0.5 0.75 0.875 0.92 0.94

Zero inlet whirl at all radii is a form of free vortex and with constant work with
radius the stator inlet flow is also free vortex. Another special case is that of
100% reaction constant with radius. As shown in figure 3.2 the stator inlet and
stator oudet whirl velocities are equal and opposite in this case and are therefore
inversely proportional to radius for constant work. This is the case where free
vortex and constant reaction occur simultaneously; it is not commonly adopted in
modem designs. A feature of free vortex is that the axial velocity is constant
with radius when the total pressure is radially constant

Constant reaction

Whereas free vortex designs have the advantage of simplicity in that with
radially constant total pressure, they satisfy simple radial equilibrium with
radially constant axial velocity, this is not so for radially constant reaction,
except in the special case of 100% reaction, which also satisfies the free vortex
conditions. Some designs of several decades ago used constant reaction and
assumed constant axial velocity, ignoring radial equilibrium. It can be shown
that, after a few stages, the flow will tend towards the design assumption.
However this would not be contemplated in modem design practice; at least
some form of simple radial equilibrium would be adopted, but with the aid of
computerised calculation, full equilibrium, including the influence of meridional
streamline curvature, is commonly allowed for.
The use of free vortex has the disadvantage that the Mach number relative to
the rotor inlet increases rapidly with increasing radius. Before the development
of transonic blade design, it was generally considered desirable not to exceed a
design Mach number of 0.8 or so. This was a severe limitation to free vortex
designs and made a choice of vortex giving a more nearly constant radial
distribution of Mach number attractive. The assumption of radially constant axial
velocity and constant reaction of 50% moves in this direction, but of course the
flow departs from equilibrium. The assumption of radially constant total and
static pressures is a first step towards radial equilibrium. As can be seen in figure
8.1 the total velocity vector Vo must increase with radius when Va is assumed
constant with 50% reaction. Constant total and static pressures imply Vo is
constant and Va reduces somewhat with radius since Va = Vo.cosao, and oto
must increase with radius to give constant work. These assumptions also give a
radially constant Mach number at rotor inlet, as discussed in Chapter 2.
58 Axialflow fans and compressors

The assumption of simple radial equilibrium, which assumes the meridional


streamlines lie on cylindrical surfaces concentric with the axis of rotation,
requires the static pressure to increase towards the outer casing, since this
provides the centripetal force necessary according to the equation:

dp/dr = pVw*/r

where p is the air density. This results in Vo falling with increasing radius and
ao increases more rapidly with radius for constant work; Va therefore falls more
rapidly towards the casing. These effects can be beneficial in terms of the Mach
number distribution and blade loadings so far as rotor inlet conditions are
concerned.

Outer radius Mid radius


(Common diagram)
Constant axial velocity
Constant Vo and constant static pressure
------------ Simple radial equilibrium with change of Va
across the blade rows.
Figure 8.1 50% reaction diagrams for various vortex assumptions
At stator inlet the whirl velocities are significantly greater than at rotor inlet, as
are the flow angles a 3 compared to cxo. The result is that the gradient of axial
velocity required for simple radial equilibrium is very much steeper than at rotor
inlet for 50% reaction. Indeed, if the diameter ratio is low the axial velocity at
the casing can drop to zero. This is obviously impractical, and constant reaction
designs assuming simple equilibrium at both rotor and stator inlet should have a
reaction of not less than 60%. The increasing gradient of the axial velocity
distribution at stator inlet necessitates an inward movement of the streamlines
between rotor inlet and stator inlet. There is a corresponding outward movement
from stator inlet to the following rotor inlet. These movements imply meridional
curvature of the streamlines. If the blade axial chords are large in relation to the
radial movements of the streamlines the static pressure gradients required to
balance the meridional curvatures will be small, and simple radial equilibrium
will give a close approximation to the flow. This will tend to be the case for high
diameter ratio, low aspect ratio blading, such as found at the rear stages of multi­
Vortexflow 59

stage compressors. For early stages the opposite tends to be the case. The
extreme example is to assume the rotor has negligible axial chord, which is
referred to as an actuator disc. In this case there can be no change of axial
velocity within the rotor, although the change of whirl velocity is assumed to
occur as in a rotor of finite chord. The theory of the actuator disc is fully
discussed in Horlock (1978). The flow may be considered to be in simple radial
equilibrium far upstream of the rotor and also far downstream. At the plane of
the actuator disc approximately half of the final downstream radial redistribution
will have taken place. The axial velocity distribution will therefore be
intermediate between the upstream and downstream distributions.
When a stator follows closely behind the rotor the simple equilibrium
condition is not achieved at either blade exit. Instead the stator tends to induce a
radial distribution midway between the rotor and stator simple equilibrium
conditions. This is because the mean whirl velocities within the stator are the
same as the mean whirl velocities, in stationary co-ordinates, within the rotor.
This suggests that a suitable approximation for low diameter ratio, high aspect
ratio stages is to assume the mean whirl velocities are in simple radial
equilibrium. Measurements in a low diameter ratio, lowspeed research
compressor tended to confirm this approximation. At the first rotor inlet the flow
was close to simple radial equilibrium, but at the first stator inlet the
redistribution was only about half that required for simple equilibrium. Further
downstream in the two stage compressor the distribution of axial velocity
changed little across the blade rows.
For a constant reaction design with constant work radially we have:

Vw0 = (1-R)U - AH/2U

and: Vw3= (l-R)U + AH/2U

Hence: Vwffl = (l-R)U

For simple equilibrium of the mean whirl velocity:

dp/dr = pVwra2/r

and writing: K = 1-R; and U = cor, gives:

dp/dr = p K V r /r

/dp = /K2pto2rdr

Pr2 - Pri = 0.5p(l-R )V (r22-r,2)


60 Axialflow fans and compressors

For constant total pressure of the vector mean velocities:


Pr2 -Pn = 0.5p(Vnv,2 - Vm*22)

Hence: Vttvi2 - Vm^2 = (l-R)2co2(r22-ri2)

By substituting Va2 + Vwa2 for Vm2 and then (1 - R)2U2 for Vwro2 in the above
expression a relationship for the axial velocity is obtained as:

Va22 = Va,2 - 2(1 - R)2(U22 - U,2)

Arbitrary vortex

Constant reaction is not the only alternative to free vortex. Designers have
chosen a variety of arrangements such as constant cto or constant a 3. Some
voitex arrangements are chosen because they allow simpler blade design e.g..
constant section stators; others because they provide some preferred radial
distribution of Mach number or loading. In all cases it is highly desirable to carry
out axisymmetric through flow calculations taking account of the full
equilibrium equations. The output from these computations must be carefully
examined to ensure that the axial velocity ratio across any blade section does not
depart too far from unity. Reasonable limits are 0.85 to 1.15. Outside these limits
the choice of suitable blade geometry may become problematical. The de Haller
number can also be examined as a first indication of blade loading, as well as
deflection angles and Mach numbers.

Simple equilibrium

To balance the centrifugal force of a particle rotating in a circular path with a


radial pressure gradient it can be readily shown that the pressure gradient
required is given by:
dp/dr = pVw2/r

Assuming the total pressure and the density to be radially constant we can write:

p = P - V4pV2

and hence: dp/dV = -pV

where V is the total velocity vector. Hence the pressure gradient can also be
written as:

dp/dr = (dp/dV)(dV/dr) = - pVdV/dr


Vortexflow 61

Hence: VdV/dr = - Vw2/r

Also: Vw = Vsina

So: VdV/dr = - V2sin2a/r

and: dV/V = - (sin2a/r)dr

Which integrates to give:

lofeV,, - lofeV,, =
i r

Only if sina can be expressed analytically as a function of r can the RHS be


integrated directly; usually the integration must be earned out numerically, in the
case where sina is constant we obtain:

V/V0 = (ro/r)"2®

Examples of simple equilibrium

Figure 8.2 shows the radial distributions of axial velocity corresponding to


simple radial equilibrium for radially constant angles of 15°, 30° and 45°. In
these cases the variation of axial velocity is due entirely to the variation of static
pressure. This increases with the level of the whirl velocity component and
accounts for the increasing negative gradient of axial velocity as the whirl angle
increases. Figure 8.3 shows the simple equilibrium axial velocities for two linear
distributions of the whirl angle. In these cases the axial velocity varies due to the
variation of the angle as well as the variation of the static pressure. The variation
of the axial velocity for radially constant static pressure is indicated by the
curves labelled ‘Const.Vo’. These illustrate that the effects of static pressure
variation and angle variation are of the same order in these cases, which are
similar to those required for constant reaction. Where a variable stagger guide
vane or stator has an angle distribution similar to the 0° to 30° graph, the 15° to
45° angle distribution would correspond approximately to the case for a 15°
increase in stagger. The resulting increasing gradient of the axial velocity
distribution means that the reduction of incidence on to the following rotor is
much greater at the hub and less at the tip than at the mean radius. Much the
same can be said of the data of figure 8.2. A falling axial velocity towards the
outer diameter has the advantage that the radial distribution of loading
parameters can be more favourable than for a constant axial velocity.
62 Axialflow fans and compressors

0.6
0.4 0.5 0.6 0.7 0.8 0.9
Radius ratio r/r*

Figure &2 Simple redial equilibrium axial velocities for constant angles
* cto- 15° to 45°

1 0.8

I 07

0.5 * ■Const
» » -Vo» • ■ • ■
0.4 0.5 0.6 0.7 0.8 0.9
Radius ratio r/r*

0.4 0.5 0.6 0.7 0.8 0.9


Radius ratio r/r tip

Figure 8 J Axial velocity profiles for linear angle variations

Considered for the same axial velocity at the mean radius, the increased axial
velocity at the hub increases the de Haller number there and conversely at the
casing. A constant axial velocity with radius tends to produce an over loaded hub
and an under loaded rotor tip when radially constant pressure ratio is required
and for this reason single stage transonic fans for high by-pass ratio turbofan
engines are often designed with an increasing pressure ratio towards the casing.
Vortexflaw 63

The use of a suitable distribution of inlet whirl angle can therefore offer
loading advantages as well as a reduction of the rotor inlet relative Mach
number, particularly when a radially constant pressure ratio is required. Whether
to employ inlet whirl must be balanced against the penalties entailed in the
provision of the additional blade row required. Where variable whirl is required
an inlet guide vane is essential, and although a zero cambered vane is possible,
the advantages of inlet whirl at design speed tend to indicate a cambered vane
producing an appropriate radial distribution of inlet whirl velocity.

Constant ‘reaction* design

It should be noted that if simple radial equilibrium exists between each blade
row there cannot be radially constant reaction. This is because the static pressure
gradient at rotor inlet is less than at rotor outlet due to the increased level of
whirl velocity at rotor outlet. The static pressure rise across the rotor tip must
therefore be greater than across the hub section, and for radially constant total
pressure rise the reaction must increase towards the tip. The only exception is a
design for 100% reaction at all radii, where the whirl velocities at rotor inlet and
outlet are of the same magnitude but opposite sign.
A design for R = 0.6 and obeying the equations:

Vwo « (l-R)U - AH/2U


and: Vw3 = (l-R)U + AH/2U

may be called 60% reaction, but is not strictly so for the reasons outlined above.
The axial velocity distributions for this design, having a work coefficient of 1.0
at the hub are given on figure 8.4. The vector mean axial velocities are also
indicated.

Radius ratio r/r ^

Figure 8.4 Axial velocities for typical ‘constant reaction’ design


64 Axialflow fans and compressors

The air angles for the design of figure 8.4 are shown on figure 8.5. Note that
these indicate a negative deflection at the rotor tip, i.e. a t is less than a*. The
static pressure rise across the rotor tip is generated entirely by streamtube
expansion and axial velocity reduction rather than deflection to a lower relative
air direction.

Radius ratio r/r,

Figure 8.5 Air angles for 60% ‘constant reaction’ design


with simple radial equilibrium

IGV Rotor Stator Rotor Stator

Axis of rotation

Figure 8.6 Meridional streamline curvature

Assuming the streamlines of figure 8.6 are such as to satisfy the axial velocity
profiles required for simple radial equilibrium, it is clear that there are significant
meridional curvatures at the inter-blade row planes, except between the IGV and
the first rotor, where the streamlines have points of inflection as the curvatures
change from concave inwards to concave outwards. The curvatures at all other
inter-blade row gaps are such as to require a radial gradient of static pressure of
negative sign at rotor outlet planes, and positive at rotor inlets. The resulting
pressure gradient will therefore be less than the simple radial equilibrium
gradient at rotor outlet planes and greater than the simple radial equilibrium
gradient at rotor inlet planes. This will reduce the differences in the axial
velocity distributions at the two planes towards the distribution for simple radial
Vortexflow 65

equilibrium of the vector mean velocities. As mentioned previously this indicates


that a good approximation to the real flow is given by simple radial equilibrium
of the flow between the IGV and first rotor and thereafter of the vector mean
velocities. Figure 8.7 illustrates the axial velocities for simple equilibrium of the
vector mean conditions for various values of reaction from 50% to 100%. These
are independent of the level of the work coefficient, but are dependent on the
flow coefficient, which was set at a value of 0.7 at the mid-radius for these
examples. It will be noted that at 50% reaction the axial velocity falls to zero just
inside the tip radius. For lower flow coefficients this tendency is more maiked.

Radius ratio r/r,

Figure 8.7 Axial velocities for simple equilibrium of vector mean velocities

Throughflow computations

As indicated earlier a throughflow computer program is commonly used to


calculate the radial distributions of all the required parameters, and this accounts
for the true equilibrium of the flow on an axisymmetric basis. However, it is
necessary to make appropriate inputs to such a program and unless the designer
has a good understanding of the basic concepts involved, it could be a very
tedious process to arrive at a suitable design by trial and error. Even with such an
understanding, a number of trials are likely to be required to achieve the best
compromise for all parameters at all radii.
9 Mach number effects

Introduction

The Mach number is the ratio of the local velocity to the local velocity of sound.
When the Mach number is unity pressure waves cannot be propagated upstream
in the flow and the nature of the flow is significandy altered. Even when the
upstream Mach number is as low as 0.5 local velocities within a blade passage
may reach the local speed of sound and modify the flow. While industrial fans
generally operate at low enough velocities for Mach number effects to be
negligible, gas turbine and industrial compressors operate at high subsonic or
even low supersonic velocities and are designed with attention to the influence of
the Mach numbers involved.

High subsonic Mach number

As the relative inlet Mach number to a blade row increases the range of
incidence over which the total pressure loss is a minimum reduces at both
positive and negative incidences. The loss loops, as shown on figure 9.1 become
narrower, but the minimum loss remains the same to a Mach number of 0.5 in
this case. Above this the minimum loss also increases and the loops continue to
narrow. For any inlet angle the Mach number at which the loss is twice the low
Mach number minimum is known as the critical Mach number. The Mach
number giving a minimum loss of twice the low Mach number minimum is
known as the maximum critical Mach number. At an incidence which becomes
less negative as Mach number increases the loss loop becomes vertical and this
defines the maximum Mach number v incidence. Figure 9.2 shows plots of
critical and maximum Mach number against incidence for a typical cascade.
Clearly it is preferable to place the desired design point within the loop formed
by the critical Mach number curves. The exact position may depend on whether
a larger margin from choke or stall is desired. To avoid premature stall at low
Mach number effects 67

speeds a front stage may be designed nearer choke than stall, while for the same
reason a rear stage may be designed nearer stall in order to avoid premature
choke at low speeds.

Max critical Mach I

0.08 Mn - 0.4 0.5 0.6 0.7 0.8

0.06

0.04

I 0.02

0
0 10 20 30 40 50 60
Inlet angle oi

Figure 9.1 Loss loops at various Inlet Mach Numbers

0.3 0.4 0.5 0.6 0.7 0.8


Inlet Mach No.

Figure 9 J Critical and maximum Mach number plot for typical cascade

Passage throat width

A throat or minimum passage width can be determined for typical cascades. As


indicated in figure 9.3 this is taken as the diameter of the largest circle which can
roll through the passage, denoted *0*. If the velocity across the throat width were
constant the blade passage would choke when this velocity was equal to the
velocity of sound i.e. Mach number = 1.0. Because of the curvature of the blade
surfaces and the surface boundary layers the velocity varies across the throat.
68 Axial fans and compressors

Figure 93 Blade passage throat width

however the concept of a constant velocity provides a basis for a theoretical


maximum Mach number for comparison with the experimental value determined
from cascade testing. The inlet flow width, delivering to the throat, is Scosat and
so the inlet Mach number would equal the throat Mach number when 0/Scosat
= 1.0. When ot| is less than the value giving this result the theoretical maximum
Mach number is less than 1.0 and the inlet flow accelerates to choke the throat

0.6 0.7 0.8 0.9 1 1.1 1.2


O/Scosctt

Figure 9.4 Theoretical and actual maximum Mach numbers

When 0/Scosot| is greater than 1.0 the throat width is greater than the approach
width. The increasing flow width from inlet to the throat could cause a
supersonic acceleration, as occurs downstream of the throat in a
convergent/divergent nozzle, in which case the throat Mach number would be
supersonic. This does not normally occur in a cascade tunnel designed for
subsonic operation, and the experimental maximum Mach number does not reach
1.0 as shown in figure 9.4, where the theoretical and experimental maximum
Mach numbers are compared against O/Scosoti.
Mach number effects 69

It has been shown that when the inlet angle is 60° the maximum critical Mach
number occurs when the ratio O/Scoscti is 1.15 and at 45° it falls to 1.0. This
trend is similar even at very low Mach number as was shown by McKenzie
(1980). One of the most important aspects of blade design at high Mach number
is to ensure that the throat width is not choked at the desired operating condition.
If the design is carried out for low speed, and then checked to show that there is
at least a 5% throat width margin from choke at the design Mach number,
satisfactory performance should result
A knowledge of the throat width of cascades is obviously essential to the
design process at high Mach number. It can be obtained by drawing, but this is a
laborious procedure. Modem computer graphics methods could be used for this
purpose, where available. A number of graphical correlations have been
produced, based on the normal blade geometry parameters. One of these
methods, for DCA blades, is given by Wright and Miller (1991). The same
paper gives a simple relationship for the variation of the minimum loss inlet
angle with Mach number as:

O/Scoscti = 0.155M] + 0.935

Supersonic Mach number

Few cascade tests have been conducted at supersonic inlet Mach number because
of the complexity of the tunnel geometry required. Neither has it been possible
to operate stator blades successfully in compressors at supersonic inlet Mach
numbers. The same is not true of rotor blades, however, and the so called
transonic compressor is a feature of many modem aero engines. In the 1950's
several attempts were made to design fully supersonic compressors. In the main
these were unsuccessful, but it emerged that the rotors in general gave a superior
performance to the stators if the latter had a supersonic inlet Mach number. This
led to experiments with transonic designs, which feature rotors with a supersonic
Mach number at the tip, falling to subsonic at the hub. The stators generally have
subsonic Mach numbers.
The most obvious reason for the superior performance of rotors at supersonic
inlet Mach number compared to stators is the centrifuging effect on the blade
boundary layers in a rotor blade, which is absent in stators. Boundary layer
particles in a rotor are accelerated to a whirl velocity close to the blade speed as
their relative velocity to the rotating blade falls. The radial static pressure
gradient is determined by the main stream flow and is less than required for
radial equilibrium of the boundary layer, which therefore tends to migrate to a
greater radius. This migration leads to additional work being done on the
boundary layer which re-energises it. Due to the considerable growth of the
boundary layer thickness which accompanies a passage shock the centrifuging
effect is of considerable advantage. Its absence in stator blades may be a major
reason for the relatively poor performance of stator blades at supersonic inlet
70 Axialfans and cofnpressors

Mach number. Another factor which may also operate in the same direction is
the fact that a supersonic relative Mach number at rotor inlet is achieved with a
subsonic Mach number relative to the casing, which is not the case for a stator.
This may mean much greater losses in the casing boundary layer and additional
secondary losses for a supersonic stator compared to a supersonic rotor.

Blade forms for supersonic operation

Initially the blade profiles for transonic rotor blades were of DCA form, and
these were quite successful to the extent that they are still considered a suitable
choice for Mach numbers up to 1.2. The highest Mach numbers employed are
about 1.6 at the tip section of high by-pass ratio fans. While DCA profiles were
initially used for this purpose, modem developments of computer generated
aerofoils giving more efficient shock structures have now superseded them. For
supersonic operation it is generally desirable to have the inlet air angle to the
blade approximately parallel to the blade upper surface. A relatively sharp
leading edge is desirable and a compromise must be struck between the
aerodynamic requirements and the mechanical requirements, especially with
regard to foreign object damage, large birds being a particular problem for
turbofan aero engines.

Passage shock waves

When operating supersonically a shock wave forms in the blade passage. Across
the shock there is an unavoidable loss of total pressure. This is accompanied by a
rise of static pressure and the Mach number drops from supersonic to subsonic.
The flow is in principle similar to the divergent portion of a conveigent/divergent
nozzle. Appendix A of Cohen, Rogers, and Saravanamuttoo (1972) gives a
useful summary of the phenomena, and demonstrates that the efficiency of the
compression of static pressure is in excess of 90% up to an approach Mach
number of 1.5, resulting in a static pressure ratio of almost 2.5:1. The efficiency
falls to less than 80% at a Mach number of 2.0, at which the static pressure ratio
is 4.5:1. The losses due purely to the shock are increased by the rapid growth of
the boundary layer which inevitably occurs downstream.
At low pressure ratios the shock occurs towards the trailing edge of the blade.
As the back pressure is increased at constant speed the shock moves forward in
the passage while the incidence remains constant. This results in an increasing
pressure ratio at constant mass flow, i.e. constant or unique incidence. When the
shock is just behind the leading edge on the lower surface of the blade the
efficiency reaches a maximum. Further increase of the back pressure expels the
shock which becomes a detached bow shock. The mass flow reduces rapidly
with further increase of back pressure giving a rapidly flattening pressure ratio v
mass flow curve and the efficiency falls off. Figure 9.5 illustrates the shock
Mach number effects 71

patterns and the corresponding positions on the pressure ratio and efficiency
characteristics.
Relative
flow
A- Detached
bow shock

ui Normal
shock near
leading edge

Mb m flow % d K lgn C- Normal


shock near
trailing edge
Operation from C to B It at the *unique' incidence

Figure 9.5 Shock positions and operating points


for supersonic rotor blade section

S/C requirement

An important design feature for supersonic operation is the need for rather lower
S/C than for a subsonic blade of similar deflection. This is due to the subsonic
boundary layer on the blade upper surface ahead of the shock. The boundary
layer being subsonic, the sudden rise of static pressure across the shock in the
main stream cannot be developed in the boundary layer. A complex flow pattern
with a rapid growth of boundary layer thickness takes place and the pressure rise
occurs over a streamwise distance of considerable extent, as indicated on figure
9.6.
On the pressure surface almost the whole blade chord is available for this
process to take place. The pressures on both surfaces have to equalise at the
trailing edge, and to provide an adequate length of surface behind the shock on
the suction surface the S/C has to be relatively low. It is found that there is a
remarkable similarity of all the major manufacturers high by-pass fans in this
respect. At the tip sections of the rotor blades the stagger angle is about 60° and
the S/C is close to 0.8.
Although transonic stages with inlet relative Mach numbers to the rotor tip up
to 1.6 have shown acceptable overall fan efficiencies in the mid to high eighties
the radial distribution of efficiency shows a considerable fall off towards the
outer diameter as indicated in figure 9.7. Perhaps surprisingly, this trend does not
diminish rapidly as rotational speed and relative Mach number are reduced. This
72 Axialfans and compressors

may indicate that it is due as much to secondary flow losses as to the high Mach
number at the design speed.

Length required for Inadequate length for


subsonic diffusion boundary layer diffusion
Figure 9.6 Requirement for low S/C with high stagger supersonic blading

0 10 2 0 3 0 4 0 50 60 70 80 80 100
0 1 0 2 ) 3 0 4 0 9 ) 6 0 7 0 8 0 9 0 100
%*■«*■M0I

Figure 9.7 Radial distribution of pressure ratio and efficiency


for a transonic fan rotor
After Nicholas and Freeman (1982) with permission of AdAA and ICAS

Laser anemometry

The development of transonic blading has been greatly assisted by the use of
laser anemometry. This non-intrusive type of instrumentation allows the
velocities within the rotor blade passages to be determined, and in particular the
position of the shocks. Comparison of these measurements with computational
predictions of the blade to blade flow field has proved a powerful method of
improving the predictions and hence of arriving at more satisfactory blade
designs at much less cost than time consuming ad hoc development.
Mach number effects 73

Effect of Mach number on overall pressure ratio

The effect of increasing Mach number on the pressure ratio which can be
produced by a single stage is illustrated by figure 9.8. A stage of typical work
and flow coefficients at a modest polytropic efficiency and zero inlet whirl has
been assumed for this diagram.

Rotor Cp»- 0.5

0.5 1 1.5
Rotor inlet relative Mech No.

Figure 9.8 Pressure ratio of a single stage with increasing Mach number

The pressure ratio available from ten stages of constant mean radius design is
illustrated by figure 9.9. In this case a 50% reaction stage is assumed at 88%
polytropic efficiency and a de Haller No. of 0.707 (Cpi = 0.5). A pressure ratio
of 7 is indicated at a first stage inlet relative Mach number close to 0.7, which
would be representative of the state of the art about 1950. A pressure ratio of 25
is indicated at a Mach number of 1.0, which is more typical of present day
technology.

40

0
0.4 0.6 0.8 1 1.2
Stage 1 rotor relative Inlet
Mach No.

Figure 9.9 Pressure ratio available from 10 stages


74 Axial fans and compressors

It is clear that increasing design Mach number has been the major contributor
to increased pressure ratio from a given number of stages. This has required the
development of improved aerofoil sections and particularly lower thickness
chord ratios. The latter have only been possible because of the use of lower
aspect ratio blading together with titanium alloys in place of steel or aluminium
alloys in the early stages. The consequence has been a great reduction in the total
number of blades required for a given pressure ratio. Because of the lower aspect
ratio of the blades the overall length of the compressor is not reduced in
proportion to the number of stages.
10 Reynolds number effects

Introduction

The Reynolds number is a non-dimensional parameter which indicates the ratio


of the inertial forces of the fluid flow to the viscous forces. A non-dimensional
group can be derived by dimensional analysis as described by Shepherd (1956)
and gives:

Re = pV//n

where Re is the Reynolds number, p is the fluid density, V the velocity, / is a


characteristic linear dimension, and p. is the fluid viscosity. The viscosity and
density are often combined as li/p = »»which is the kinematic viscosity.
Strictly, only geometrically similar bodies can be compared by their respective
Reynolds numbers and then any linear dimension can be used. However, it is
still found useful to compare machines which are 'similar* in a looser definition,
and it is conventional for axial compressors to quote the Reynolds number based
on the inlet gas conditions, the velocity relative to the inlet of the first stage rotor
at mid-height, and the rotor blade chord, also at mid-height. This assumes that
the variation of chords from blade row to blade row is not dramatically different
from one compressor to another.
While the effects of Mach number are dominant for high performance
compressors, the effects of Reynolds number are still significant if the value
should be changed by a factor of 1.5 or more, which will typically cause a
change of one percentage point of efficiency. For industrial fans and other low
speed machines Mach numbers may be so low as to be insignificant and
Reynolds number is the more important parameter.
76 Axialflow fans and compressors

Performance variation

It has been indicated by Carter et al. (1957) that the variation of losses due to
change of Reynolds number can be described by the equation:

l-H = k R e D

For Re > 0.5*105,the power n = - 0.2

For Re < OJMC^.the power n = - 0.5

Bullock (1964) quotes a similar equation:

(l-T l,)/(l-il 2) = a + b(Re1/Re2)c

He points out however that in the literature the values of ‘a’ vary from 0 to 0.5,
of *b* from 0.5 to 1 and of *c’ from 0.1 to 0.2. He concludes a detailed
discussion with the observation that c = - 0.2 is satisfactory for Reynolds
numbers greater than 4*10* provided the blade surfaces are aerodynamically
smooth. He also indicates that the unsteady pressure and velocity fields produced
in compressors by blades passing through the wakes of preceding blade rows are
probably responsible for the comparatively low Reynolds number at which the
effect of turbulent flow is evident in the behaviour with varying Reynolds
number as compared to cascade tests.

WassePs correlation

Wassell (1967) uses the same equation as Carter et al, but puts the index n =
p*q, where p is determined by the level of mean Mach number through the
compressor, and q is determined by the ratio of the ’effective’ length to the mean
annulus height He develops an impressive correlation of data from twenty
different compressors which includes the effects on surge pressure ratio and
mass flow as well as efficiency. The correlation is limited to Reynolds numbers
greater than the critical, but this is indicated to be as low as 0.25* 105 for single
stage fans and multi-stage compressors, compared to 1.0*10* for cascades.
Because it is one of the most practically useful methods for Reynolds number
corrections the essentials of Wassell’s correlation are given in figures 10.1
through 10.3. In figure 10.1 for the parameter p the base has been changed from
V/VT to mean Mach number in order to avoid any confusion over units. The
mean Mach number can be taken as the value at the design speed relative to the
inlet of the mid-stage rotor, or the average of first and last stage values. All other
parameters are explained on the diagrams.
Reynolds number effects 77

Application of the correlation

Use of the correlation is best illustrated by an example. Say a compressor has


been tested at an inlet pressure of 60 kPa and the results are to be corrected to an
inlet pressure of 101.3 kPa. The inlet temperature is 288 K and the rpm constant
for both cases.

OGV Mean annul us h t ■ ‘/j [hi+h?l


h. ha U , - UN/(N - 7x)
N * No. of stages

Efftctlv* length I iman annulus


Mmui Mach No. height

Figure 10.1 Wassel efficiency correction parameters


Adaptedfrom Wassel (1967) with permission of ASME

Efficiency correction

For the efficiency correlation Wassel defines a Reynolds Number Rei based on
the mean relative inlet velocity to the first stage rotor, the first stage rotor chord
and the inlet density and viscosity. If this Reynolds number is 0.7* 105 at 60 kPa
inlet pressure it will be 0.7(101.3/60)* 103 = 1.18*105 at the higher pressure,
since density is proportional to pressure at constant temperature. For the
efficiency correlation let the mean Mach number be 0.7, which gives p = 0.85.
Also let the length of the rotor ( L r ) be 400 mm and the mean annulus height 109
mm. The effective length (1^) is given by:

= 400*6/(6 - 0.5) = 436 for the 6 stages.

The correction is to allow for the additional length of a typical stator as


compared to the long chord or double row outlet guide vanes used at the outlet of
some compressors. The mean annulus height is 109 mm, which results in a value
78 Axialflow fans and compressors

of 4 for the ratio of effective length to mean annulus height, and from figure
10.1(b) the value of q is 0.14, and hence n = p*q = 0.119.
For the lower Reynolds number 1 - Tio.7 = W.TMO5)"0119

For the higher one: 1 - Hu* = k( 1.18* 10s)]*0119

Hence we can write: 1 - n u i = (1 - tjXO.7/1.18)0'119

Thus, if Ho.7 = 85%, then: 1 - n,.,t = 0.15(0J93)°119, and n,.„ = 85.9%

Mass flow correction

For the mass flow correction the linear dimension of the Reynolds number, Re2,
is defined as the distance x from the first rotor leading edge to the point where
the throat width of the passage intersects the suction surface of the blade, as
shown on figure 10.2.

Re* * piVj„*x/Hi
Reynold** No. R** 10*4

Figure 10J Wassel mass flow correction


Adaptedfrom Wassel (1967) with permission cfASME

If x/C = 0.5, then Re2 = 0.5Rei and so Re2 = 0.35* 105. From figure 10.2, if Olx
= 0.9 then:

(0/*XQa/Qa* - 1 ) 52-0.05

Hence Qa/Qa* = 1 - (0.05/0.9) = 0.944

At standard atmospheric inlet pressure the Reynolds number Re2 will be :

Rej = OJRe, = 0.5* 1.18* 105= 0.59'K)3

Hence: (O/xXQa/Qa*- 1) = -0.025


Reynolds number effects 79

and: Qa/Qa* = 1 - (0.025/0.9) = 0.972

Hence Qs*/Q«o = 0.972/0.944 = 1.03

Thus the mass flow is estimated to be 3% greater at standard atmospheric inlet


pressure compared with the test pressure.

Surge pressure ratio correction

For the surge pressure ratio correlation another Reynolds number Re3 is defined

as: Re3 = piVah/^i

where Va is the axial velocity at inlet to the first stage rotor and h is the first
stage rotor height For typical geometry this results in Re3 = U Rei

Reynold's No. Re»* 10**


Figure 103 Wassel surge pressure ratio correction
Adaptedfrom Wassel(1967) with permission of ASME

For Re, = 0.7* 105, Rej = 1.05M05 and for Re, = 1.18*10*. Re3 = 1.77M05

At Rej = 1.05M03figure 10.3 gives (Rs - Rs*)/Rs* = 0.01

and at Re3 = J.77‘ 105: (Rs - Rs*)/Rs* = 0.05

Hence RsWRsw = 1.05/1.01 = 1.04

which results in a corrected surge pressure ratio of 5.2, for a test value of 5.
80 Axialflow fans and compressors

Surface finish effects

It is well known that the fhction factor for pipe flow becomes constant above a
value of the Reynolds number which varies with the surface roughness of the
pipe. The value of the friction factor falls as the roughness is reduced and the
Reynolds number at which it becomes constant increases, as described by
Moody (1944), and adapted by Shepherd (1956). Similar effects are to be
expected in compressors with variation of the blading surface finish. Schaffler
(1979) presents the results of a series of compressor tests which agree reasonably
well with Wassell's correlation but indicate a constant efficiency above an upper
critical Reynolds number which is a function of the surface finish of the blades.

Velocity
Turbulent boundary layer
Roughness within laminar sublayer
* hydraulically smooth
'Laminar sublayer
Roughness protrudes Into turbulent
layer * hydraulically rough
Figure 10.4 Aerodynamic roughness

Approx. Reynolds number

Figure 10.5 Roughness effect on efficiency


Reynolds number effects 81

The mechanism, which is shown diagrammatically in figure 10.4, is that when


the roughness is such that the peaks of the surface irregularities do not protrude
through the laminar sub-boundary layer the turbulent boundary layer is
unaffected and the flow behaves as if the surface is aerodynamically smooth. At
some higher Reynolds number, implying a relatively thinner boundary layer, the
peaks of the surface irregularities will protrude into the turbulent boundary layer
and cause the flow to behave as if the surface is aerodynamically rough,
producing a constant efficiency with further increase of Reynolds number.
Hence the increase of efficiency with increasing Reynolds number is dependent
on the quality of the blade surface finish. Schaffler's results illustrate clearly that
polished blade surfaces, particularly on the rear stages of a compressor, will
delay performance deterioration due to falling Reynolds number at increased
altitude.
Miller (1977) quotes the upper critical Reynolds number as:

R„c =16C/kd,

where C is the blade chord and kd» is the surface roughness measured by the
Centre Line Average method, which is quoted in micro-inches i.e. inches* 1C* =
25.4 nm
For forged blading a CLA value of 32 is representative, and if polished this can
be reduced to 16, with the effect shown in figure 10.5. Care should be taken to
ensure that the Reynolds number is high enough to exceed the upper critical
before incurring the cost of polishing blading. This may lead to polishing only a
number of rear stages and ignoring forward stages where the Reynolds number is
less than the upper critical over the operating range. For small compressors it
may be found that polishing to a CLA value below 32 is ineffective in producing
any improvement, whereas for large turbofans it can produce worthwhile
improvements, but not requiring polishing of the front half of the core
compressor stages.
It has been suggested that polished blade surfaces may reduce the amount of
dirt which is accumulated on the blades over a period in service and that this
effect may justify polishing of blades where the above arguments would not
necessarily justify the cost involved. It could also be that polishing to a smoother
surface finish is justified for this reason.
11 Compressible flow
relationships

Introduction

It is convenient both for the calculation of flow properties in ducts and for the
non-dimensional presentation of compressor performance characteristics to use a
number of nondimensional and quasi-nondimensional groups. These can be
developed from a few fundamental relationships.

Notation

a = local speed of sound = CyGt)1/2 m/s


A = local cross sectional flow area m2
Cp = specific heat at constant pressure kJ/kgK
Cy = specific heat at constant volume kJ/kgK
Y= Cp/Cv
G = gas constant = 0.287 for air kJ/kgK
M = mass flow rate kg/s
Mn= Mach Number = V/a
p = static pressure kPa
P = total pressure kPa
t = static temperature K
T = total temperature K
V = local velocity m/s
p * density kg/m3

Fundamental Relationships

Gas Law p/p = Gt d)


Compressible flow relationships 83

Continuity of mass flow M = pAV........ ..........................(2)

Energy T « t + V2/2Cp................................. (3)

Isentropic relation Ftp = (TA)y(T“!)...............................(4)

G = Cp-C*........................................(5)

From (3) and (5) we can obtain: TA a 1 + 0.5(y- 1)Mh2.........................(6)

and from (4) P/jp = {1 + 0.5(y - OMn2}**"0................... (7)

Rearranging (3) V2/2Cp = T -t;

hence V/V(CpT) = {2(1 - 1fT))in.......................(8)

From (1) and (2) M/A = pV/Gt

hence: Q = MVT/AP = (1/GXV/VTXTAXp/P).............. .(9)

and q = MVT/Ap = (1/GXV/VTXTA)..................... (10)

From equation (8) it is apparent that V/>/(CpT) is non dimensional and hence Q
and q are only quasi-nondimensional The true nondimensional forms are:

(MVT/APXGWC,) = [VMCpT)](TAXp/P)

and (MVT/ApXG/VC,) = [V/V(CpT)](TA)

When written in this form any self consistent set of units will give the same
values. It is common practice, however, to use the quasi-nondimensional forms
given by equations (9) and (10). Since the value of Cp varies appreciably for the
range of temperatures encountered for air compressors, different tables or curves
are required for various ranges of temperature. Gases other than air will also
require their own values for the relationships.
Since: G = Cp - C*; and y = Cp/Cy. we have: Cp = G{*y/(Y - 1)}. While G is
constant at 287 kJ/kgK for air in the range of temperature of interest to
compressors, the value of Cp rises with increasing temperature and it is usual to
specify the data by the value of y which falls from about 1.4 at 250K to 1.35 at
850K. A typical set of graphs are shown on figure 11.1 plotted to a base of Mach
number. The most notable feature of these curves is that Q reaches a maximum
when the Mach number = 1.0, while all the other parameters increase
continuously. An excellent resume of the theory of gas dynamics which explains
the background to this and other transonic phenomena is given in an appendix to
84 Axialflow fans and compressors

‘Gas turbine theory' by Cohen et al (1972). A comprehensive set of tables for


isentropic compressible flow and related topics are given by Palmer et al (1987).

Applications

Airflow measurement

By means of a well flared entry to a ducting system the mass flow rate may be
measured by means of total and static pressures as indicated in figure 11.2. By
measuring the dynamic pressure, i.e. total - static, and the total as a difference to
atmospheric, which should be very small if the flow is drawn directly from
atmosphere, the ratio of the absolute values of total and static pressures are
derived. The value of Q is then obtained from tables or graphs. Provided an
effective flow area for the duct cross section at the measuring plane is known
and the atmospheric temperature has been measured the actual mass flow rate
may be calculated:

M = (MVT/AP) (AP/VT)

The effective flow area divided by the geometric area is known as the
discharge coefficient, Q , and can be determined by measurements of total
pressure across the duct boundary layer in the plane of measurement For a well
designed inlet flare the value of C<j will be about 0.99.

Mach No.

Figure 11.1 Compressible flow relationships for air (y = 1.4)


Compressible flow relationships 85

Static tapping
Airflow ^ Pitot tube

B * Barometric pressure
P * Total pressure
p ■ Static pressure

Figure 11.2 Airflow pressure measurements

Duct total pressure loss

The loss of total pressure between two planes in a ducting system may be
obtained in the following manner. Assuming the mass flow rate is known, a
number of static taps may be positioned on the duct wall in the plane of interest
This should be well clear of bends or major obstructions in the duct so that the
static pressure can be assumed constant across the plane. Total temperature will
be constant along the duct. Qi is known for the upstream plane as in the
preceding paragraph, hence:

<b = Qi(Al/A2)P,/pj and P2/P2 = /( qi); so P2 = PzCPi/Pi)

and hence the pressure loss Pt - P2 can be obtained.


It should be noted that only if the velocity is constant across the measuring
planes will the total pressure derived in this way be in agreement with the value
derived from an area traverse of total pressure. The latter will give a larger value
than that derived as above from static pressure and continuity. The difference
will depend on the amount by which the velocity varies and the method of
averaging the total pressure measurements. Three methods of averaging the total
pressure are:
• Mass weighting
• Momentum averaging
• Area averaging

These will give descending values of total pressure in that order. For further
details on this subject the reader should consult Livesey and Hugh (1966).
86 Axialflow fans and compressors

Compressor discharge total pressure

In a similar manner to the above the total pressure at outlet of a compressor may
be obtained from measurement of the static pressure. In this case the outlet total
temperature must also be measured.

cb=Qi(p./p2W a y r 1xA1/Aj)

P2/P2 = /(q 2)

P2 = PldPrfpl)

The remarks made previously about variation of the velocity obviously apply to
this situation as well. An effective area may be assumed, but this is likely to vary
with the compressor operating condition.

Ductflow properties

To find the duct area necessary to pass a flow of lOOkg/s at a MN = 0.25 when
the total pressure = 405.2kPa and the total temperature = 457K.

From flow data at Mn = 0.25 Q = MVT/AP = 16.8

M>/T/P = 100*^457/405.2

A = (M>/T/P)/(M>/T/AP) = 5.276/16.8 = 0.314m2

To find the mass average velocity in the duct.

From flow data V/VT = 5. V = 5.0V457 = 106.9 m/s

To find the static pressure:


P/p = 1.044 at MN= 0.25

p = P/(P/p) = 405.2/1.044 = 388.1kPa

To find the dynamic pressure in mm Hg.

Dynamic pressure = P - p = 405.2 - 388.1 = 17.1kPa

To convert this dynamic pressure to the equivalent of Mercury column:

101.3kPa = 760 mm Hg, hence:- 17.1kPa = 760*17.1/101.3 = 128.3 mm Hg


12 The stage characteristic

Introduction

The performance of an individual stage can be represented approximately as a


unique set of curves for all rotational speeds, in terms of AH/U2, and AP/(pU2)
plotted to a base of Va/U, as previously shown on figure 1.7. This is usually
called the stage characteristic. It is only genuinely unique with variation of
rotational speed for a limited range of Reynolds number and when the blade inlet
relative Mach numbers are low, say less than 0.3. As Mach number increases
some variation appears in the curves obtained due to falling efficiency and to
varying axial velocity ratio across the blading, resulting from increasing density
ratio as speed increases. These variations are sufficiently small that the concept
of the unique stage characteristic is an extremely useful one. By measurement of
casing static pressures between blade rows individual stage or blade row
characteristics can be derived from multistage tests and provide a useful analysis
tool for performance development Prediction of individual stage characteristics
provides the basis for stage stacking performance prediction methods as
discussed in Chapter 22.

The work coefficient characteristic

The basic parameters of the stage characteristic are the work coefficient AH/U2,
and the flow coefficient Va/U. As discussed previously, if the air oudet angle
from the blades were to be constant independent of the incidence, AH/U2 would
vary linearly with Va/U. Although the outlet angle is nearly constant at
incidences well away from stall, it rises significantly near stall and as indicated
by Howell (1945), the deflection may reach a maximum at stall. At that point the
outlet angle must rise at the same rate as the incidence. Despite this, most stage
characteristics tend to indicate a work v flow coefficient curve which is
approximately a straight line above the peak efficiency flow coefficient although
of a lower negative gradient than corresponding to constant outlet angles. Howell
and Bonham (1951) gave graphical data for the slope of the curve at the design
88 Axialflow fans and compressors

point. This indicated that the gradient of the curve diminished as the work
coefficient increased. Data derived from the experiments reported in McKenzie
(1980) tended to show that Howell and Bonham's correlation could be improved
if the S/C was introduced as a parameter. Intuitively it can be readily appreciated
that the greater the spacing of the blades the greater will be the tendency for the
outlet angle to increase with incidence.

Figure 12.1 The work characteristic slope.


Based on Howell and Bonham (1951)

Writing y for AH/U2 and yd for the design value and referring to figure 12.1
where <p is the flow coefficient, Va/U, <pd is the design value, and tan©' is the
gradient of the work coefficient characteristic for constant outlet angles and tan@
is the actual characteristic slope at maximum efficiency, then:

tan 0 / = (l/\|/d) -1

tan 0 = {(l/\jfd) - 1} {1 - 0.4S/C}

The constant 0.4 is for circular arc camber line blades. It may be smaller, say
0.2 perhaps, for blades with less camber towards the trailing edge such as
controlled diffusion profiles.

The efficiency characteristic

Howell and Bonham also give a curve for the relative efficiency, Tl/rim**, in terms
of the relative value of AH/UVa, which will be written x^Xd- This is reproduced
in figure 12.2 together with a modified version derived from a variety of sources.
The stage characteristic 89

x/u
Figure 1Z2 Relative efficiency correlation
Adaptedfrom Howell and Bonham (1951) with permission of 1 Mech E

The difference between the two curves at low values of %/%&is thought to be due
to moderately high Mach number data being incorporated, whereas the Howell
and Bonham data is thought to have been based on low speed data only.

The pressure rise coefficient

This is the stage characteristic equivalent of the pressure ratio. There are three
possible ways of expressing it One is r|AH/U2 and a second is derived from this:

t|AH/U2 = nCpAT/U2 = TiCp(AT/r,)/(u7r,)

and riAT/T, = R<r'yi' -1

Hence tlAH/U2 = C,Ti {R<r‘1^ - 1)/U2

The third form is appropriate for low Mach numbers only where the density
change across the stage is negligible. It can be derived from the above form by
assuming the pressure rise is very small in proportion to the inlet pressure.

When AP/P, is small R ^ 1* - 1 -((y-D/ylAP/P,.

As shown previously (y -1 )ly = G/Cp

and so ((y-l)/Y)AP/P = (G/Cp)AP/(pGT,) = AP/fpCpT,)


90 Axialflaw fans and compressors

Hence the pressure rise coefficient can be written:

t)AH/UJ = AP/(pU2)

This latter form can be used to high speeds if a mean density is used to allow for
the variation of density across the stage.

Experimental stage characteristics

Normally only the pressure rise and flow coefficients are derived from
multistage tests. Figure 12.3 shows typical examples for a first, middle, and last
stage. For the first stage the points obtained at 50% of design speed are at low
flow coefficient and on a positive gradient, indicating the stage is stalled. At
75% design speed they are around the peak of the curve, while at design speed
the points are to the right of the peak on a negative gradient At 110% of design
speed the points are at lower pressure rise coefficient than design speed and
indicate a near vertical gradient These points lie on a line which is at lower flow
coefficient than an extension of the curve through the design speed points. These
are indications of choking of the first stage at 110% of design speed. Compared
to the first stage, the middle stage shows the points for all speeds grouped much
more closely together. They do not spread far to the left of the maximum
pressure rise or far below the design point on the choke side. Thus the range of
flow coefficient required of a middle stage is less than for a first stage. This has
sometimes been interpreted to mean that the design loading parameter (Cp*, Dp
ctc.) may be higher on a middle stage than an early or late stage, since such a
large stall margin may not be required. The final stage characteristic shows no
operation in the stall region. The highest speed points are at high pressure rise
and low speed points are at low pressure rise and high flow coefficient Clearly
the rear stage is limiting the maximum flow of the compressor at low speeds and
the maximum pressure ratio at high speeds.

Experimental measurement

One of the most commonly used methods of obtaining some knowledge of


individual stage performance from tests of multi-stage axial compressors is by
measuring outer casing static pressures between blade rows. By invoking a few
simplifying assumptions these measurements, together with the corresponding
overall performance measurements of mass flow, rotational speed, and
temperature rise, can be expressed in the form of pressure rise and flow
coefficients. Study of the resulting plots for the individual stages, or blade rows,
can give a valuable insight to the stage matching.
In the simplest arrangement static pressure tappings are placed in the stator
casing between the blade rows. It is desirable to use four or five tappings at each
axial station. To obtain a good average pressure it is desirable to have a number
The stage characteristic 91

0.5 First stage 0.5 Mid stage 0.5 Laatstage

0.4 -I— i— A O

0

o
0.4 0.4 - x>
\


b ► X
i 0*3 X
< X
0.3 % 0.3 . %
02 X
X %
0.1 -> 02 -h ,i ... 1 1 0.2
0.3 0.4 0.5 0.6 0.7 03 0.3 0.4 0.5 0.6 0.3 0.4 0.5 0.6 0.7
Va/U Va/U Va/U

x 110% Nde. o 100% N*. A 50% N*.


X 110% Nde, O 100% Nde. A 50% Nde*

Figure 123 Typical stage characteristics (4:1 design pressure ratio)

of tappings at different circumferential positions relative to the stator blade pitch,


but in front of different blade passages.
If individual blade row performance is required it is particularly important to
have multiple tappings suitably spaced ahead of the stator leading edges. When
variable stator stagger is employed, it is more desirable to analyse individual
rotor and stator blade rows.

Temperature measurement

In general it is not practical to measure temperature at every stage, and some


analysis methods rely on proportioning the overall temperature measurement
between the stages. One method is to divide the measured overall temperature
rise in proportion to the stage design temperature rises. Another is to calculate
them on the assumption that each stage has a polytropic efficiency equal to the
overall value. Neither of these methods are satisfactory for high pressure ratio
machines when operating far from their design points, and give rise to much
scatter of the stage characteristic data for the middle stages in particular.
Even for a design pressure ratio as low as 4:1 a marked improvement can be
made by measuring temperature at one middle stage. For higher design pressure
ratios two or three intermediate temperature measurements are desirable.

Analysis method

For each compressor operating point the following data is required:

M = mass flow rate, kg/s


Tj = compressor inlet total temperature, K
P, = compressor inlet total pressure, kPa
92 Axialflow fans and compressors

?2 = compressor outlet total pressure, kPa


T2 = compressor outlet total temperature, K
An = annulus area at entry to n,h rotor row, m2
Am = annulus area at entry to stator row, m2
Tn = total temperature measured at i)* stage stator, K
Ooo = air exit angle from ny, stator (pitch line design value), degrees
(X2a = rel. air exit angle from 11* rotor (pitch line design value), degrees.
Pm = average static pressure at entry to n<h rotor, kPa
pn = average static pressure at entry to nth stator, kPa
Va = mass average axial velocity, m/s
U = blade speed at mid-span of each rotor, m/s

Suppose temperatures are measured at compressor inlet and at thefourthstage


stator. If the design temperature rises of these four stages are 15, 20, 25,and 25
degrees respectively then of the temperature rise measured over the four stages
(15/85)*ATobi would be the temperature rise of the first stage, and (20/85)*ATob.
for the second stage. Alternatively all the four stages can be assumed to operate
with the same polytropic efficiency. Thus:

(iWp,.)(1"l)ft" = T rf/ r 1

from which rj may be derived, and hence:

T „ /r1= (iWp,i)(’H)'n7

and: TV T, = (p,3/Pr1)(T“‘VnT and T^/T, = etc.

Further improvement of the accuracy of the stage temperature rises could be


obtained by iteration of the flow coefficient and an approximation to the work
coefficient characteristic.
Having determined the total temperatures, the flow coefficient at any rotor inlet
plane can be calculated.

M V T W A n , COSCtofo-i)Pm = ( M V T j / P |) ( P i / p mX T ,( n - i) /T i) 1/2/A n , COSCto(n-l)

Hence from compressible flow relationships:

VhlT<»n=M
where q is the LHS of the previous equation. The flow coefficient is given by:

Va/U - (V/VT,(I>.|))‘N/(T,(n.j))cosao(n.i)/U
The stage characteristic 93

The blade speed, U, is usually the mid span value, but can be the hub or dp
value if preferred. The various forms in which the pressure rise coefficient can
be expressed have been discussed above. Since the velocities at exit from
succeeding stators are generally very similar the static pressure rise and total
pressure rise are very similar for the stage as a whole. It is simplest therefore to
use the static pressures directly to express the pressure rise. Where it is desired
to use total pressures, these can be obtained from compressible flow
relationships giving P/p as a function of V/VT as determined for the flow
coefficient
Where it is desired to derive characteristics for rotor and stator blade rows
separately the static pressures can again be used directly. It must be borne in
mind that since the whirl velocities are generally greater after rotors than stators,
the radial static pressure gradient is also greater after rotors. This results in the
rotor static pressure rise on the casing being greater than at the mean or hub
radius. For the stator the casing pressure rise is less than at the mean or hub
radius. The size of these differences is dependant on the diameter ratio and the
stage reaction. For high reactions and diameter ratios the effects are small, but
increase for low reactions and low diameter ratios. Since the blade row
characteristics are often used only to indicate whether the blade is operating
stalled or choked it is the gradient of the curves which are of greatest concern
rather than the value of the pressure rise. Unless corrections are made to allow
for the different radial static pressure gradients after rotors and stators it is
dubious whether it is justifiable to calculate the separate blade row pressure rise
coefficients based on total pressures, especially as the calculation of the rotor
exit total pressure is more complex.

VaAJ

Figure 12.4 Stage characteristics for variable stagger stators

Analysis of the stage characteristics where variable stagger stators are


employed is no different except that the appropriate value of the stator outlet
94 Axialflaw fans and compressors

angles must be used. This is assumed to be the design value increased or


decreased by the change from design stagger. The stage characteristics for
variable stagger stators can be particularly useful in determining the optimum
settings for the variable vanes of each stage. Some typical stage characteristic
plots for a stage with variable geometry are shown on figure 12.4.
13 The repeating stage concept

Introduction

It is well established that after a few similar stages the flow develops a radial
distribution which repeats after every stage. This is particularly obvious in terms
of the radial distribution of axial velocity, but is also true for other parameters
such as flow angles or total pressure. The initial effects of casing and hub
friction in the first few blade rows appears to rapidly reach an equilibrium
condition similar to fully developed pipe flow. Wherever the total pressure losses
are high the axial velocity is reduced and the work done is increased and vice
versa. The result is to maintain a radially constant increment of total and static
pressure across the stage.
For multi-stage low speed research compressors with identical stages, it has
been found that the flow approaches the repeating stage condition by the third
stage. Four stage research machines have been adopted in a number of cases for
this reason. The fourth stage makes the third a 'buried* stage, and avoids the
study stage suffering from any special influences occurring at the final stage.

The linear repeating stage

The repeating stage concept is useful in discussion of the radial matching of high
diameter ratio stages in particular. Since the diameter ratio is usually high
towards the rear of multi-stage compressors the repeating stage condition is most
nearly true there. In the following discussion the diameter ratio is assumed to
approach unity, or the blades are assumed to move linearly instead of rotating.
Thus we can introduce the term 'linear repeating stage*. This allows two
assumptions to be made which greatly simplify the argument. These are that both
the blade speed and the static pressure are constant along the blade span. In such
a compressor, stages having constant section rotors and stators would have a
'design* condition of spanwise constant axial velocity and work.
96 Axialflow fans and compressors

%Span
Figure 13.1 Typical spanwise profiles for a repeating stage

For such a stage the work coefficient v flow coefficient would be a unique
characteristic for all points along the span, provided any influence of secondary
flows or wall boundary layers on deviation are ignored. The pressure rise
characteristic would also be unique if the efficiency were only a function of the
flow coefficient, and did not depend on the spanwise position. However, it must
be accepted that there will be more loss close to the annulus walls, and the
efficiency at a given flow coefficient will fall towards the walls. For some
distance either side of the mid-span the profile loss will be the only cause of
inefficiency, but closer to the walls secondary losses will cause a progressively
lower efficiency. At the mid-span the efficiency, for modestly loaded blade
sections, will lie in the region of 93% to 95%, while the overall stage efficiency
is likely to be 88% to 90%. The maximum efficiency achieved at the walls is
indicated to be in the region of 70% to 75%, see for example Lakshminarayana
et al (1994). Radial profiles of axial velocity and efficiency, such as sketched in
figure 13.1, can be found which will determine a variation of work along the
span such that when the mass integral of the work is calculated, the mass average
stage efficiency has a typical value, and the total pressure rise is radially
constant
The breakdown of the stage inefficiency into annulus friction, secondary, and
profile loss elements suggested by Howell (1945) is shown in figure 13.2. The
summation of the annulus and secondary losses are indicated to cause an
approximately constant inefficiency over a wide flow range of the stage
characteristic. This suggests that the efficiency at any given spanwise position
may be a constant amount less than the midspan efficiency, which is determined
by the profile losses alone.
If the spanwise distribution of efficiency can be determined at one operating
point, then efficiency curves can be constructed over the whole flow range on
this basis. The pressure rise curves for corresponding spanwise positions can
The repeating stage concept 97

Va/U

Figure 13.2 Annulus, secondary, and total losses.


After Howell(1945) with permission o fl Mech E

Va/U

Figure 133 Operating points for a linear repeating stage

then be constructed using the unique work characteristic. Figure 13.3 illustrates
the resulting curves for efficiency, work and pressure The pressure rise can be
either the static or total value, as the repeating stage requires these to be the
same. Operation of the stage at a given mass flow requires the pressure rise to be
the same all along the span. Points A, B and C represent one operating condition,
and the distribution of axial velocity is determined by these points. The work is
significantly greater near the walls than at midspan, and a mean work coefficient
can be found from:
AH„/U2= J AH/UJd(Va/U)/Jd(Va/U)

It should be noted that the enthalpy or temperature distribution, according to


this procedure, will not repeat from stage to stage, but will continually increase
98 Axialflow fans and compressors

more at the walls than at midspan. In practice it is found that the spanwise
variation of temperature is limited at the rear stages, and research studies by
Adkins and Smith (1982) and Gallimore and Cumpsty (1986) have indicated that
a considerable amount of spanwise mixing takes place, thus reducing the
temperature gradients. A first approximation to allow for this in the design
process is to accept the variation of work in the rotor according to the vector
diagram and to assume that the temperature variation mixes out completely in
flowing through the stator.
The pressure rise characteristics of figure 13.3 indicate that the maximum stage
pressure rise will be limited by the section characteristic adjacent to the wall,
since this has the lowest maximum value and the pressure rise must be the same
at all points along the span.

Stage of unity work coefficient

An interesting result follows when the procedure is applied to a ‘linear’ stage


having AH/U2 =1.0. For such a stage AH/U2 is constant and independent of the
flow coefficient, assuming the air outlet angles are constant The efficiency has
the same value as the pressure rise coefficient since AP/(pU2) = T|AH/U2
As indicated on figure 13.4 this results in the efficiency at a given operating
point being constant along the span. Also the maximum efficiency is limited to
the maximum achieved at the wall. This contrasts with the more typical stage of
figure 13.3 where all the sections are reasonably close to their maximum values
simultaneously. It is reasonable to expect that if a reduction of efficiency does
occur when AH/U2 = 1.0 it will appear progressively as unity is approached.
Some evidence to support this is available from the data reported in McKenzie
(1980) and from tests conducted for Rolls Royce in the water compressor
described by Howell and Bonham (1951), and is shown in figure 13.5. Ruffles
(1991) has also indicated that efficiency tends to fall with increasing design
AH/U2, as shown by figure 13.6.
It must be noted that at AH/U2 = 1.0 the flow coefficient is likely to be in
excess of unity and therefore the relative motion effects discussed in Chapter 7
are not so favourable as at lower values of the work and flow coefficients, where
it is more conventional to design. In practice, even at the highest diameter ratio
of about 0.9, normally allowed at the rear stage of a high pressure ratio
compressor, the work coefficient at hub and casing will be in the ratio of 0.92 =
0.81 and so the work characteristic will vary considerably in slope along the
blade span. The constant spanwise efficiency derived from the linear stage
argument does not therefore arise in practical stages. It remains however that the
trends of falling efficiency with rising design work coefficient shown in figures
13.5 and 13.6, may be due to a greater degree of spanwise mismatching brought
about by the generally lower slope of the work characteristics.
The repeating stage concept

i.i
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
Va/U

Figure 13.4 Linear stage with AH/U2= 1

100 McKenzie (1980) data*


Water Compressor °
% ♦
90

o<o
80
UJ ♦ □

70 ■ ■ » ‘ ‘ * ‘
0 0.2 0.4 0.6 0.8
AH/U*

Figure 13.5 Trend of peak efficiency with work coefficient

£
£

Mean stage loading AH/U*

Figure 13.6 Polytropic efficiency trends


Adapted from Ruffles (1991)
14 Spanwise matching of high
diameter ratio stages

Introduction

A basic approach to this subject has been described in discussion of the repeating
stage concept in Chapter 13. The objective of this chapter is to examine what
design action may be possible to improve the radial matching and thus the
maximum efficiency and range of the stage performance. The repeating linear
stage concept is used in order to simplify the discussion. Application to the
design of practical stages will not present a problem when the basic concepts are
appreciated.

Variable axial profile

The deterioration of the axial velocity distribution through the early stages, to the
equivalent of fully developed pipe flow, led to consideration of improved
performance by adjustment of the blade geometry near the end walls. Andrews et
al.(1956) described two designs which attempted to account for the end effects.
One design was known as Variable Axial Profile (VAP), and the other as
Variable Work Done (VWD). In the VAP design the blade geometry was
modified so as to optimise the incidences to suit the angles measured in the
datum design, while the design work was maintained as for the datum. This led
to changes in the design vector diagram as indicated in figure 14.1 for a station
close to the casing. At the mid-span the VAP diagram had a higher axial velocity
than the datum. The result of tests on the VAP blading showed a loss of
efficiency and a deterioration of the velocity profile such that the departure from
the design intent was of the same order as for the datum design. This result can
be qualitatively explained using the repeating linear stage concept with the
assumptions made in discussion concerning it. The pressure rise and efficiency
characteristics for a blade section near an end wall are shown in figure 14.2. It is
clear that due to the lower efficiency the axial velocity drops below the design
Spanwise matching o f high diameter ratio stages 101

Figure 14.1 Variable axial profile and datum vector diagrams


Des.effic1!
i - e f f i c W,
0.4 0.9
.......
Operating^
ef f i ci ency| or l

Pres.Coeftr^*| 0.5
I Operating Va/tf' 0.3
0.2 Design Va/U — +\
0.1
•* Design intent
— Actual
0.1 - 0.1
0.3 0.4 0.5 0.6
Va/U

Figure 14.2 Operation of VAP stage at a section near an annulus wall

value to give the design pressure rise. At the mid span, where the efficiency will
be greater than the design value, the axial velocity will also be greater than the
design value. There will therefore be an increased variation of the axial velocity
along the span compared to the design intent and so the test result is explained in
this respect. Examination of loading parameters near the end walls, such as ideal
pressure recovery factor, diffusion factor, and deflection, shows they are all
significantly increased compared with the original design, and further increased
by the departure of the operating conditions from the design intent. This may
well explain the loss of efficiency from the datum design.

Variable work done design

The concept of the VWD was to increase the design work towards the end walls,
with the objective of maintaining the original design axial velocities across the
span. A comparison of the datum and VWD vector geometry at a station near an
102 Axialflow fans and compressors

end wall is given in figure 14.3. Section characteristics for three spanwise
stations are shown on figure 14.4, using the same assumptions as previously. The
design objective appears to have been achieved in that the axial velocity is
constant along the span at the design pressure rise. The experiment showed this
to be true to some degree; the efficiency, however, was again lower than for the
datum blading. Although there is an increase of design woik towards the blade
ends, there is no increase of actual work, and the velocities relative to the blades
at inlet are increased. Loading parameters are not increased to the same extent as

Figure 143 VWD and datum vector diagrams for a


station near an annulus wall

VaAJ

Figure 14.4 Operation of VWD linear repeating stage

for VAP. A partial explanation for the loss of efficiency may lie in the mass
integration effect. The increased axial velocity towards the end walls where
efficiency is low, and vice versa at mid-span, leads to an increase of mass
integrated work at a given pressure rise, and hence a loss of efficiency. This can
typically be of the order of 1%, and so offers at least a partial explanation of the
2 % deficit reported by Andrews et al.
Spanwise matching of high diameter ratio stages 103

Figure 145 Constant Vi inlet vector diagrams for near wall stations

An alternative approach

These qualitatively successful explanations of the VAP and VWD experiments


give some confidence in this approximate approach to an understanding of high
diameter ratio stage performance. It is therefore of interest to examine the
possibility of improving performance using this method of analysis. No attempt
is made to increase the maximum efficiency at any spanwise station, only to
optimise the matching so that all sections are closer to their maximum
efficiencies simultaneously. The influence of mass integration is allowed for and
local overloading of the blading is to be avoided. A typical efficiency for a
conventional stage design would be 90% at low Mach number and high
Reynolds number. Of the ten points of inefficiency four or five would be due to
profile losses and the remainder to tip clearance, annulus friction, and secondary
flows. The most that can therefore be reasonably expected is about 2%
improvement from the present analysis. The increment which can be caused by
mass distribution is therefore important This indicates that the axial velocity
should be designed to be as low as possible towards the end walls. The limit will
be the deflection or other loading parameter limitation. Since the static pressure
rise is equal at all points along the span, the actual pressure recovery factor Cp
would also be constant if the relative inlet velocity were constant This suggests
that compared to a conventional design the vector diagram should be modified as
indicated in figure 14.5 for a section near the end wall. The increase in design
work is as required by the spanwise efficiency variation. The relative inlet
velocities are maintained constant, and the inlet vectors are rotated to achieve the
required increase of work. The outlet velocity vectors are reduced by this change
to the vector diagram. This represents an increase to the theoretical static
pressure recovery factor, but not to the actual pressure recovery factor. The
diagram therefore attempts to represent the true local efficiency, while the
conventional datum diagram assumes a constant spanwise efficiency.
104 Axialflow fans and compressors

When (he sum of <Xo and ot| is less than 90° then <Xo is reduced by the
modification, and cto is increased when the sum is greater than 90°. As discussed
in Chapter 7, when Oo + (X| = 90° the inlet dynamic pressure to the following
blade remains substantially constant as the axial velocity varies. It would appear
that vector diagrams of this type may also go some way to allowing for end
effects automatically. Although the incidence would be increased if the blade
were not modified, it should be recalled that a hypothesis was described in
Chapter 7 which suggests that a greater incidence may be tolerable when large
secondary losses are present. It may be that these considerations offer a possible
explanation for the difficulty which has been experienced in attempting to
significantly improve the performance of conventional stage designs by attention
to end effects.
On the other hand it is possible that the relatively poor performance often
found for stages of low Oo+ai may be improved significantly by such attention.
This was illustrated by the example of a stage where AH/U2 = 1.0 in Chapter 13.
This may indicate that stages of relatively high work coefficient could respond
more readily to end treatment, at least to raise their efficiency to the level of
more conventional stages. This could provide a useful increment in stage
pressure ratio, particularly where blade speed is limited, for example with core
booster stages on the same shaft as a high bypass ratio fan. It could also be
helpful for compressors pumping gases of high specific heat and sonic velocity
such as helium.
15 Spanwise matching of low
diameter ratio stages

Introduction

When the diameter ratio is low the vector diagram and the blade geometry differ
radically from hub to casing. Assuming constant work along the span at the
design point we have: AH/U2 = constant/r2. Thus for a diameter ratio of 0.5,
AH/U2 will change by a ratio of four to one between hub and casing. The lowest
diameter ratio used in practice is about 1/3, for which the ratio of AH/U2 at hub
and casing will be nine to one. The off design problem this creates is due to the
change of slope which the different radial values of AH/U2 cause in the pressure
rise-mass flow characteristics at different radii.

Example

Let the vector diagram at the area mean radius be such that Va/Um = 0.694 and
AH/Um2 = 0.274 with zero whirl at inlet to succeeding rotors at all radii. If the
diameter ratio is 0.4, the area mean radius is given by:-

rm2 - rh2 = r,2 - r m2


which reduces to:
r™/r, = (0.5(l+(rh/rl)2))1'3

For Th/r, = 0.4, this gives TjTt = 0.7616.

To give constant work and axial velocity with radius at the design point we can
derive values as tabulated in table 15.1
106 Axialflow fans and compressors

Table 15.1

r/r, 1.0 0.945 0.828 0.688 0.515 0.4


(V a/U )k>c 0.529 0.559 0.638 0.768 1.026 1.321
( A H /U 2) * 0 .1 5 9 0 . 1 7 8 0 .2 3 2 0 .3 3 6 0 .5 9 9 1 .0

In the first line the four intermediate radii represent the mid radii of four equal
area annuli. The subscripts k* indicate that the coefficients are based on the local
radii.
V aA Jioc = (V a/U m X rm /rioc)

A H /U 2* , = ( A H /U m2) ( r J r io c ) 2

The performance curves for each radial section of the blading can be estimated
by one of the conventional methods, such as given in Chapter 22, or in Appendix
II to Howell (1945). For convenience the pressure rise is related to the mean
blade speed rather than the local blade speed of the section concerned. Similarly,
the flow coefficient is expressed as local axial velocity to mean blade speed
Vate/U*.

Plane 1 Radially constant velocity, total and static pressures


Plane 2 Radially constant total pressure only
Plane 3 Radially constant static pressure only
Figure 15.1 Model stage with shrouds
The characteristics derived above are two dimensional, and assume a constant
axial velocity across the stage. An approximate model which is helpful to an
understanding of the off-design performance of low diameter ratio stages is to
assume shrouds in both rotor and stator separating the four radial sections as
indicated in figure 15.1. The shrouds ensure that the mean axial velocity is
constant across any section from rotor inlet to stator outlet. Any adjustment of
Spanwise matching of low diameter ratio stages 107

the axial velocity distribution to satisfy downstream conditions must therefore


take place upstream of the rotor, as indicated in figure 15.1.
Some distance upstream of the first rotor the total and static pressures can be
taken to be radially constant As the flow approaches the rotor the total pressure
will remain constant across the annulus, but the static pressure may vary as the
flow starts to adjust itself to the downstream conditions. Immediately
downstream of the stage the static pressure can be assumed radially constant,
which is the condition for simple equilibrium where there is no whirl velocity.
Each radial section of the stage must therefore operate from the same inlet total
pressure to a common outlet static pressure. By subtracting !/2(Va/Um)2 from the
total to total pressure rise coefficient the characteristics for the various sections
can be converted to the form (p2 - Pi)/(pUm2). In this form the operating points
for a given overall mass flow will all lie on a horizontal line representing a
common value of this form of the pressure rise coefficient on figure 15.2. It is
clear that the inner section has the lowest maximum pressure rise, and hence this
will be the limiting value for the stage as a whole according to the assumptions
made. An average flow coefficient for the stage can be derived as:

(Va/U)AV= (Va/Url + Va/Urt + etc)/4

Va*jU.

Figure 15.2 Stage characteristics for four radial stations

This can be evaluated for a series of arbitrary values of the pressure rise
coefficient The conventional total pressure rise coefficient is given by:-

AP/pUn2 = (pi - P,)/pUmJ + 0.5(Va/Um)2AV.

For an identical stage following the first, the static pressure will be radially
constant at the second stator trailing edge, and therefore the static pressure rise
across the second stage will be equal at all radial stations. If the axial velocity
distribution is assumed to be the same at inlet and outlet of the second stage the
108 Axialflow fans and compressors

total pressure rise will be equal to the static pressure rise and the performance of
the second stage can be estimated in the same way as for the first stage but using
total to total pressure rise characteristics in place of total to static characteristics.
The radial distribution of axial velocity resulting at a given mean flow
coefficient can be compared with the distribution derived for the same flow
coefficient for the first stage. Where the differences are small the performance
prediction will be satisfactory. Where differences of axial velocity or local flow
coefficient are greater than say 5% for the same flow a more complicated
procedure is required. This entails allowing for the radial distribution of total
pressure leaving the first stage and using the total to static pressure rise
characteristics with allowance for the radial differences of total pressure to
achieve the required radially constant static pressure at the second stage outlet

Va/U,

Figure 153 Predicted and test performance of low diameter ratio stages

The average flow and pressure rise characteristic is derived as for the first
stage, and both are plotted in figure 15.3 and compared with experimental
results. The overall predicted characteristic is also compared to the experimental
two stage result on figure 15.3. Considering the assumptions made, the predicted
characteristics are in good agreement with the test measurements. In particular,
the significantly lower stall pressure rise of the second stage is well predicted.

Increased diameter ratio

These effects are fundamentally due to the lower slope of the hub pressure rise
characteristics compared with those at greater radii. If the hub sections of the
blading were removed to give stages of higher diameter ratio, the stalling
pressure rise of both stages would be predicted to increase; the second by more
than the first Test results for the case where the hub diameter was increased to
Spanwise matching of low diameter ratio stages 109

0.5 Dia^atio
hub crop

0.5 dia.
Rotor Stator ratio
tip crop

The tip crop was scaled up by 5/4


0.4Rt to fit the hub crop annulus

Figure 15.4 Annulus diagrams for 0.4 and 0.5 diameter ratio

give a diameter ratio of 0.5, for the same two stage machine as discussed above,
are shown, together with comparative predictions, in figure 15.5. Again the
predictions give a good approximation to the experimental results.
Results were also obtained with the outer blade sections removed to give 0.5
diameter ratio. The experimental maximum pressure rise was actually a little
lower than for 0.4 diameter ratio, although the predicted values for 0.4 and 0.5
were almost identical, as also shown on figure 15.5.

0.7 Hub crop


.r/rt-0 .5
0.6
r/rt - 0.4
0.5 Tip crop
AP/pU* ^ r/rt ■ 0.5
0.4

0.3
Predicted data
0.2
0.5 0.6 0.7 0.8 0.9 0.5 0.6 0.7 0.8 0.9
VaAJ VaAJ

Figure 15& Effect of hub and tip cropping on two stage performance
110 Axialflow fans and compressors

The same procedure can be applied to stages with whirl velocity at stator exit
by allowing for a radial gradient of static pressure based on the design whirl
velocity distribution.

Conclusions from the approximate method

The procedure described above is interesting, not just for the results it produces,
but for the insight it provides to the causes of off-design problems of low
diameter ratio stages. Despite the broad assumptions made, the agreement with
experimental results indicates that the primary controls on the performance are
substantially correctly modelled. It follows that those influences which are
neglected must be of secondary importance. These include the annulus boundary
layer and secondary flow effects, as well as the influence of axial velocity
changes across the blading. The major indications are that the hub sections of the
blading limit the maximum pressure rise, and this limitation is mote severe the
lower the diameter ratio. It is also more severe on the second and following
stages than the first stage. In high speed multi-stage compressors this latter effect
is offset by the rapid increase of diameter ratio through the early stages,
particularly if the casing diameter is constant over these stages.
Only the unstalled performance, at mass flows greater than corresponding to
maximum pressure rise, has been considered in the preceding discussion.
Consideration of the stalled performance is given in Chapter 16.

Alternative methods

There are two more mathematical approaches to the stage performance problem.
The first is the actuator disc, and the second the axisymmetric throughflow
calculation. In the actuator disc system a blade row is assumed to have negligible
axial chord, and is usually assumed to be near the mid chord of the real blade
row. The blade row operates without any change of axial velocity from entry to
exit, because of its zero axial chord, and in this respect is similar to the
approximations made above. However, the flow is not assumed to be in radial
equilibrium at the exit of the actuator disc, but approaches this condition
asymptotically downstream. Similarly the flow far upstream is in radial
equilibrium and takes up an intermediate distribution between upstream and
downstream equilibrium at the disc. The subject has been extensively described
by Horlock (1978).
The axisymmetric through flow calculation is more realistic and forms the
basis of many modem design systems, often combined with a form of blade to
blade flow calculation. A well known form of the axisymmetric flow calculation
has been developed by Denton (1978). The program input includes the annulus
dimensions and blade axial chords. Axial velocities in general vary across each
blade row, and radial components of velocity are included in the calculation. For
Spanwise matching o f low diameter ratio stages 111

performance prediction purposes a relationship between loss coefficient and inlet


angle can be included. Variation of the outlet angle with inlet angle variation is
also desirable, and the loss and deviation should also be functions of the axial
velocity ratio across the blade row. If desired the effect of secondary losses can
be simulated by specifying a radial variation of an additional loss coefficient
This may be zero over the centre part of the span and rise towards the blade
ends. This aspect is, of course, of increasing significance at high diameter ratio
and low aspect ratio.
16 Stall

Introduction

The term ‘stall’ is used by analogy with an aeroplane wing where, as the angle of
incidence of the airflow is increased, the lift force on the wing at first increases,
but falls sharply when a critical angle is exceeded. Similar effects occur in a
compressor blade row, and in terms of the pressure rise there is a fall off beyond
the stalling incidence, i.e. below the critical flow coefficient This may be
progressive or discontinuous, as illustrated by figure 16.1. Other than the point
of maximum pressure rise arbitrary definitions of stall, such as the point of
maximum deflection, or where the total pressure loss coefficient reaches twice
its minimum value, have been used, particularly in connection with cascade
testing.

i»i
Va/U Va/U
a) Progressive stall b) Abrupt stall

Figure 16.1 Progressive and abrupt stall


Stall 113

P2

a) Progressive stall characteristic b) Abrupt stall characteristic


(Part span stall) (Full span stall)

Figure 16.2 Typical compressor characteristics


Adaptedfrom Day et al (1977)by permission of the I Mech E
Pressure rise coefl

Va/U

Figure 16.3 Total and static pressure rise characteristics


for 0.4 diameter ratio stage
114 Axialflow fans and compressors

Criteria for progressive or abrupt stall

The total to static pressure rise coefficient of a stage is defined as:

Yts = (P2 • Pi)/pU2

where pa = stage outlet static pressure.


Pi = stage inlet total pressure
p = mean density
U = mid-span blade speed.

Day and Cumpsty (1978) showed that \|/ ts is approximately constant at a value
of 0.11 from the stalling flow coefficient almost to zero flow. This is the case
whether the stage exhibits abrupt or progressive stall and appears to be true for a
wide variety of stage designs. If a stage develops a value of \j/is greater than 0.11
before stall it will have an abrupt stall and the operating point will fall along a
throttle line as indicated in figure 16.1(b). If on the other hand the stage stalls at
Yts less than 0.11, as in figure 16.3, the total pressure rise coefficient will fall
off progressively towards zero flow. The average total pressure rise coefficient,
YtT' can be derived as:

V r r = V ts + (Va/U)2/2cos2ou

where a* is the mean stator outlet flow direction. For present purposes it is
sufficient to consider the case for cu = 0°. If the efficiency at stall is assumed to
be 88 % and the approximation is made that the value of AH/UVa at the design
point is 80% of its value at stall, a curve of design AH/U2 against Va/U can be
drawn which should correspond to the boundary between stages of progressive
and abrupt stall. That this is the case is supported by the data of figure 16.4.
Because of the approximations which have been made in deriving the curve of
figure 16.4 more data would be desirable to confirm the boundary curve. It is
worthy of note, however, that the values of Howell’s nominal S/C corresponding
to the curve are in the region of 2 at low flow coefficient and 1.2 at high flow
coefficients. Since axial compressors are generally of much lower nominal S/C it
is to be expected that they may exhibit abrupt stall, while industrial fans having
much larger S/C in general often exhibit progressive stall. For low diameter ratio
compressor stages, although the actual S/C values may not exceed 1.0 the stage
may exhibit progressive stall, and the design point lie below the boundary curve.
This is due to the design work coefficient being limited by the hub at a local
value seldom much greater than 1.0. For constant work radially this tends to limit
the midspan AH/U2 to values below the boundary curve for diameter ratios lower
than about 0.5.
Stall 115

Abrupt
0.7
0.6
0.5
0.4

0.2
0.1 0.4 Dia. ratio progressiva stal
0
0 0.2 0.4 0.6 0.8
Va/U

Figure 16.4 Demarcation between abrupt and progressive stall

Surge initiation

Where one or more stages stall progressively in a multi-stage compressor they


may not initiate surge immediately, whereas a single stage entering abrupt stall
may initiate surge as was indicated by Huppert and Benser (1953). While
progressive stall is generally associated with stages of low diameter ratio, and
abrupt stall is more common in stages of higher diameter ratio, there is also a
tendency for progressive stall to occur in stages of low design work and flow
coefficients (Va/U = 0.35 or less) even at high diameter ratio. Since diameter
ratio almost necessarily increases from first to last stage it is common for early
stages to exhibit progressive stall and rear stages to exhibit abrupt stall.
Consequently, at low speeds, when mass flow is reduced and front stages stall
first, they do not immediately cause surge. In fact approximately half the total
number of stages may have to reach stall before surge occurs at low speeds. On
the other hand, at higher speeds the rear stage approaches stall first, and because
it is likely to stall abruptly, surge may occur immediately.

Rotating stall

A feature of blade row and stage stall which has received a great deal of research
attention is the phenomenon called ‘rotating stall*. This was first reported by
Cheshire (1945) in centrifugal compressors and by Emmons et al (1955) in axial
compressors. It consists of a breakdown of the flow into sectors of stalled and
unstalled flow. The stalled sectors may have a very low axial velocity, or even a
small negative one, whereas the unstalled sectors operate at a level of axial
velocity consistent with unstalled flow. The stall cells rotate in the direction of
the blade speed, but at a lower velocity. The number of cells and their speed of
rotation may vary as the flow coefficient varies.
116 Axialflow fans and compressors

In low diameter ratio stages stall cells usually appear at the outer diameter first
as flow is reduced, and spread towards the hub with further flow reduction. In
high diameter ratio stages the stall cells often occupy the whole blade span
immediately on stall initiation, and increase in the extent of the circumference
they occupy as the flow is further reduced. Figure 16.2 shows some of the types
of stall cells observed by Day and Cumpsty in a 0.8 diameter ratio low speed
four stage research compressor.
For low diameter stages which exhibit progressive stall it can be the case that
the outlet static pressure is approximately constant at its maximum value from
the stall point to zero flow. This is illustrated on figure 16.3 together with total
pressures measured at three points across the span at the stage outlet. The static
pressure shown is the mean of inner and outer wall measurements which only
differ by a small amount The total pressures show that the inner radii reach a
maximum at a higher flow coefficient and a lower pressure rise than the outer
sections of the blading, which is consistent with the analysis of Chapter 15. The
difference between the static pressure curve and the total pressure curves is an
indication of the local dynamic pressure at stage outlet. It can be seen that the
inner radius develops a very low velocity below the stalling flow which
eventually approaches zero at the lowest flow measured. Part span rotating stall
is only initiated below the flow coefficient at which the outer section of the
blading reaches a maximum total pressure rise. As indicated by figure 16.2, in
the case of high diameter ratio stages the static pressure exhibits a discontinuity
and the static pressure is again approximately constant on the stalled branch of
the characteristic. This has led to the concept first proposed by Day, that the
unstalled segments of the annulus operate on the normal unstalled branch of the
characteristic at the same pressure rise as the stalled areas which operate at
approximately zero axial velocity. As the throttle is closed the mass flow is
reduced by a reduction of the portion of the annulus occupied by unstalled flow
while the stalled part expands. For a stage which stalls progressively the
operation is similar but the unstalled flow is at the point where the static pressure
first reaches the maximum value as flow is reduced, as indicated on figure 16.3.

Hysteresis

For stages exhibiting abrupt stall a hysteresis can occur when entering and
leaving stall. Referring to figure 16.1(b), on entering stall there is a discontinuity
as the operating point falls along a constant throttle line from the point of
maximum pressure rise onto the stalled branch of the characteristic. Further
closure of the throttle takes the operating point towards zero flow. When the
throttle is opened again the flow increases along the stalled branch to a value in
excess of the point at which it first entered the stalled branch. At a flow which,
in some cases, is approximately equal to the flow at maximum pressure rise, but
in others can be greater, there is a further discontinuity as the operating point
rises along another constant throttle line to return to the unstalled branch at a
Stall 117

flow considerably in excess of the maximum pressure rise flow. Day


demonstrated that the hysteresis became larger for stages of high design work
and flow coefficient, and tended to disappear for stages of low work and flow
coefficient
The effect of hysteresis in compressors can lead to there being two completely
separate characteristics at the same speed in the range where the first few stages
may be stalled, i.e. between 75% and 90% of design speed for design pressure
ratios in the range between 6 and 10. This is illustrated by the performance
characteristics of the first four stages of a 12 stage compressor shown on figure
16.5(a). Performance points were measured below the normal surge flow at
4000,5000 and 6000 corrected rpm.
The results were also plotted in the form of an overall pressure rise coefficient
and inlet flow coefficient This allowed the stalled part of the performance
curves to be estimated for 7000 and 7720 rpm which is the design speed, as
indicated by figure 16.6. The same was done for the overall temperature rise
curves, and from these the overall pressure ratio characteristics were derived as a
function of the outlet flow parameter as shown in figure 16.5(b). It will be seen
that at 6000 rpm and above these curves have a range of outlet flow function
where two values of pressure ratio are possible. Figure 16.7 shows the
characteristics of the front and rear groups of stages for 6000 rpm to the base of
the common flow function at the interface between them. The inlet flow function
of the rear stages is taken as equal to the outlet flow function of the front stages
and the outlet temperature of the front stages is assumed constant over the flow
range of the constant speed characteristic to allow the assumption that the rear
stages will operate at a constant inlet N/VT. Obviously two overall characteristics
are possible, as illustrated on the right of figure 16.7, and plotted against the inlet
flow function the mass flows are quite different This phenomenon was observed
on the complete 12 stage compressor, the higher mass flow characteristic was
obtained if the test speed was approached from a higher speed where the rear
stages can only match with the mass flow of the unstalled front stages. The lower
mass flow characteristic was obtained if the test speed was approached from a
lower speed where the choking flow of the rear stages was less than the stalling
flow of the front stages. Another effect of rotating stall characteristics is
illustrated by the performance of a four stage compressor operating as the low
pressure component of a two shaft gas generator of a turboshaft engine. The
stalled characteristics were measured to be approximately of constant pressure
ratio as indicated on figure 16.8. The operating line was as indicated, with the
result that from start up the compressor operated on the stalled characteristics.
However as speed increases the working line no longer intersects the stalled part
of the characteristics so that when this first occurs there is a sudden increase of
mass flow and pressure ratio without increase of speed. The result is a sudden
increase of power output and faster acceleration to maximum speed. The noise
characteristic also changed quite markedly.
118 Axialflow fans and compressors
2.2

Q. 1.8 7720
o
1 1.6 7000

] 1.4

1.2 N, - 4000

1 ■ X .. . J - , .A .

50 100 150
Inlet flow function MT^/Pi
Figure l i i a Front four stage performance based on inlet flow conditions

Outlet flow function MT^/P,


Figure 16.5b Front four stage performance based on outlet flow conditions
1.6

AH/B*
AH/U 1
12
77ao' Extrapolations
** — Extrapolations

Md 8
°-8 l-tW J------- l"tW ■"-03- ■W •0.7
05 0.4 y g Vj{j /U.3 0.6 0.7

Figure 16.6 Front four stages in flow and pressure coefficient form
Stall 119

Ri-3 * overall pressure ratio


Ri-2 » LP pressure ratio
R2.3 » HP pressure ratio

Figure 16.7 Hysteresis characteristics at part speed due to front


stages operating stalled

Stalled
characteristics

6 3

I
2 IstaJfdrop x Locus of
\ outpoint maximum
^ during accei. staled flow

30 40 50 60 70 80 90 100 110
Mass flow % max. power

Figure 16.8 LP compressor stalled operation during acceleration


Adaptedfrom McKenzie and Bayne (1975) by permission ofASME

Freeman and Dawson (1983) indicate both secondary and tertiary


characteristics are possible, the secondary branches occurring with part span
rotating stall, and the tertiary characteristic when full span rotating stall is
encountered at lower flow coefficients.

Vibration

Apart from the loss of efficiency and pressure rise which are inevitably
associated with stall, it also tends to produce blade vibrations which can, in
120 Axialflow fans and compressors

Figure 16.9 Stalled flutter

certain cases, lead to rapid fatigue failure of the blades. The fluctuating forces
which lead to the vibrations can be due to the rotating stall cells. Clearly, if the
passing frequency is close to a blade natural frequency of vibration, a resonance
will occur and large amplitudes may be produced with associated fluctuating
stresses leading to fatigue failure in a matter of seconds in the worst cases.
Another source of vibration is stall flutter. This is due to negative damping of
the lift force on the blade in stall. Referring to figure 16.9, when operating
unstalled, i.e. on the positive slope of the lift v incidence curve, if there is a
movement of the blade to the right the incidence and the lift will both increase.
The increased lift will resist the motion of the blade, and acts as a strong
damping force to blade vibration. When operating above the stalling incidence
the slope of the lift curve is negative and so the damping force is also negative.
Large amplitudes of vibration may again be set up by this mechanism. In the
days when it was common to employ high aspect ratios and aluminium alloy
blading the flutter mechanism was commonly assumed to be the primary cause
of early stage blade failures, due to stalled operation at low speeds.

Passive stall control

While the introduction of titanium alloys and lower aspect ratios in high
performance compressors have given greater tolerance to stalled operation,
nonetheless it is to be avoided if possible. Consequently, considerable efforts
Stall 121

— n n tu i

Circumferential grooves

Vane

Rotor Rotor

Blade separator Air separator

Figure 16.10 Some types of anti-stall casing treatment


Adaptedfrom McKenzie (1993) with permission of I Mech E
have been made to provide a greater stall free range of operation. Methods which
depend on variable stagger stator blades, interstage bleed, and twin shafts are
discussed in Chapter 19. Here some methods which require no adjustable
mechanisms are considered. It has been demonstrated that the stage stall point
may be delayed to a lower flow coefficient by a variety of configurations of
slots, honeycombs, and plenums fitted in the casing above the rotor blades.
Figure 16.10 illustrates some of these arrangements and further details are
given by Smith and Cumpsty (1984) and by Fujita and Takata (1984). Little
theory has been successfully applied to the phenomenon, but qualitatively it has
been suggested that the mechanism is that air flow with a high whirl velocity at
the rotor tip is trapped in the slots. The whirl energy is removed by the stationary
slot, and the air is discharged into the annulus near the leading edge of the rotor
with some additional kinetic energy remaining. This appears to prevent the
establishment of rotating stall cells until a lower flow coefficient. These devices
have been successfully applied in a number of aero gas turbine compressors.
This has usually been confined to the first stage of moderate pressure ratio
compressors in multiple shaft engines. It is accompanied by a loss of 0.5% to 1%
of efficiency at the design speed on machines with design pressure ratios
between 3 and 5. The loss of stage efficiency, where the casing treatment is
applied must therefore be in the region of 2% to 5%. This tends to preclude
application to more than one stage.
122 Axialflow fans and compressors

Industrial fan designers have developed a number of passive stall delay


devices. These are generally of larger size in comparison to the blading than
those discussed above, but in many cases offer even greater extension of the
unstalled range of flow. The configuration patented by Ivanov et al (1984) has
been demonstrated by Azimian et al (1989) to operate effectively and with
minimal loss of maximum efficiency when unstalled. Azimians experiments also
showed that the improvement of the unstalled range was much greater when no
stator followed the rotor. This would appear to be because the stator hub section
is commonly the first to stall, and the casing treatment has much less influence
on conditions at the hub.
Ziabasharhagh (1992) extended this work to higher diameter ratios and also
demonstrated that the treatment was capable of restoring the loss of pressure rise
and range of flow caused by a typically distorted inlet flow. The arrangement
investigated by Azimian et al and by Ziabasharhagh, based on the patent
drawings of Ivanov et al are illustrated in figure 16.11. The mode of operation is
thought to be similar to that of the smaller scale casing treatments described
above. As the flow tends towards stall, patches of flow near the outer annulus
develop a whirl velocity of the same order as the blade tip speed, due to the very
low axial velocity which occurs in a stall cell. This flow is centrifuged into the
casing recess over the forward portion of the rotor blade. The cambered vanes
remove the whirl velocity as the centrifuged flow passes forward, and it is
discharged into the annulus upstream of the rotor with its whirl velocity
removed. The greater effectiveness of this type of treatment compared to the
smaller scale treatments may be due to the larger proportion of the flow which

Figure 16.11 Ivanov type casing treatment


Adaptedfrom Ivanov et al (1984)
Stall 123

they can handle, and the greater effectiveness with which the recirculated flow is
discharged to the upstream annulus, i.e. the lower pressure losses imposed on the
recirculated flow. A remarkable feature of these devices is that with
approximately the forward half of the rotor blade chord exposed to the recess
there is such a very small loss of efficiency when operating unstalled.

VaAJ

Figure 16.12 Performance change of 0.7 diameter ratio fan


with Ivanov type treatment
Adaptedfrom Ziabasharhagh (1992)
17 Surge

Introduction

Surge is a violent instability of the whole compressor system during which the
flow reverses and recovers at a frequency of a few cycles per second. In terms of
the overall pressure ratio v mass flow characteristic it is most commonly and
simply explained with the aid of figures 17.1 and 17.2. The compressor delivers
from atmosphere to a large reservoir in which the pressure can only change very
slowly on account of its large volume. If a small perturbation increases the mass
flow from A to C on figure 17.2 the receiver pressure will be greater than the
pressure delivered by the compressor. This causes a greater resistance to the
flow, which therefore tends to fall. As the flow falls the pressure delivered
increases again until operation at point A is restored. A small perturbation of
mass flow to B causes the pressure delivered by the compressor to exceed the
plenum pressure, leading to an increase of the flow, and again operation at A is
restored. Thus operation on the negative gradient part of the characteristic is
inherently stable.

0 Mass flow —♦
Figure 17.1 Compressor system Figure 17.2 Stability with large plenum
Surge 125

When operation at the point of maximum pressure ratio is considered, it is


clear that a perturbation of the flow from D to E will be stable, as for A to C. On
the other hand a perturbation from D to F leaves the delivery pressure lower than
the plenum pressure which therefore tends to cause a further reduction of the
mass flow. This process leads rapidly to a complete flow reversal with the
compressor operating momentarily on a negative flow branch of its
characteristic, and the reservoir discharges in both directions until its pressure
has fallen substantially. The negative flow branch of the compressor
characteristic is steep as shown on figure 17.1. The compressor now starts to
deliver again at the high flow end of its characteristic. The delivery to the
reservoir considerably exceeds the flow through the throttle, and so the resident
mass and pressure in the reservoir builds up again until the system is once more
at the point of maximum pressure ratio, and so the cycle repeats itself. This will
continue until the outlet throttle is opened sufficiently to restore operation on the
negative gradient of the characteristic.
When the volume of the reservoir is reduced, surge occurs at an increased
frequency, due to the shorter time required to empty and refill the smaller
volume. If the reservoir is reduced to a negligible volume the diagram of figure
17.3 illustrates the system stability. It is no longer appropriate to assume
momentarily constant pressure in the reservoir when the mass flow is perturbed a
small amount. Instead, going to the other extreme, the throttle characteristic is

Figure 173 Stability with small plenum


taken to apply directly to the compressor characteristic. Figure 17.3 shows a
throttle characteristic curve passing through the compressor characteristic at
point A. When the throttle is closed slightly to intersect at point C, the flow
through the throttle reduces immediately to point B. Delivery from the
compressor falls and the plenum pressure rises so that the throttle flow increases
from B to C, and the compressor operating point simultaneously converges on
point C. When operation on the positive gradient of the compressor characteristic
126 Axialflow fans and compressors

at point D is considered, where the throttle curve has a positive gradient greater
than that of the compressor characteristic, the same argument applies and
operation would be stable. At point E, where the throttle and compressor curves
have the same slope, a small closure of the throttle will cause the operating point
to drop to F, where the compressor curve again has a smaller positive gradient
than the throttle curve. This movement is unstable in the sense that flows
between E and F are unobtainable; operation will however stabilise at point F.
This is sometimes called deep stall and can be an extremely dangerous condition
because it involves rotating stall and recirculating flows within the compressor
which can lead to rapid and extreme rises of temperature, even leading to
melting of the blade material. Greitzer (1978) developed a theory to account for
whether a compressor system would develop cyclic surge or deep stall,
depending on the relative volumes of the plenum and the compressor and the
rotational speed of the compressor.

An alternative criterion for surge

The above argument implies that stable operation is possible on the positive
gradient of the compressor characteristic so long as the gradient is less than that
of the throttle curve. This is seldom the case in practice. On the contrary, it often
appears that suige occurs while the characteristic has a significant negative
gradient The tendency for surge to occur while the pressure ratio v mass flow
characteristic has a negative gradient is more commonly observed at the higher
end of the speed range, and it is the rear stages which approach their stall points
first in this region. These stages are of higher diameter ratio than early stages,
and for this reason tend to show a sudden collapse of pressure rise at the stall
point compared to stages of lower diameter ratio, which tend to show a more
progressive stall as discussed in Chapter 16. The result is that whereas a number
of early stages may stall at low speeds before surge occurs, at higher speeds if
only one rear stage reaches stall the compressor may surge. This is indicated by
the fact that if individual stage characteristics are derived from inter-stage
measurements on a multistage compressor, front stages show operating points in
stall at lower speeds i.e. the stage characteristics have regions of positive
gradient. The rear stages, on the other hand, seldom show any regions of positive
gradient. A further point in support of this argument is that commonly the surge
line in the upper speed region closely approximates to a constant value of the last
stage flow coefficient.
An alternative explanation for surge occurring before the maximum pressure
ratio is reached is that maximum density ratio, rather than maximum pressure
ratio, may be the determining factor. Referring to figure 17.4, which shows
typical temperature ratio, efficiency, and pressure ratio curves plotted against
mass flow for a constant speed, it is noted that the maximum efficiency occurs at
a lower pressure ratio and higher mass flow than maximum pressure ratio. Also
Surge 127

it should be noted that the temperature ratio curve continues with a negative
gradient to flows below maximum pressure ratio.
Density ratio is given by:
P2/P 1 = (Pa/PiVCTa/Ti)

Efficiency Max. effic’y flow

Pressure Max. pressure


ratio ratio flow
Density
Max. density
ratio
ratio flow
Temperai
ratio

Mass flow

Figure 17.4 Relative flows for maximum efficiency,


pressure and density ratios

Row entering
plenum
Mass
flow Flow leaving
plenum
Plenum Plenum
residen resident
mass mass
Time Time

Figure 17.5 Transient operation from A to B and C to D of figure 17.4


128 Axialflow fans and compressors

At the point of maximum polytropic efficiency, pressure ratio and temperature


ratio are related by:
Pi/P, = c iy ri)w<T"1)
where 11 is the maximum efficiency and T[y/(y-1) will have a value of
approximately 3. So density ratio will be:

P2/P 1 = (Tj/Tj)2

Therefore the density ratio v mass flow curve will also have a negative gradient
at this point. At the point of maximum pressure ratio, where temperature ratio
still has a negative gradient, the density ratio will have a positive gradient since a
small positive increment of mass flow will cause a drop of temperature ratio but
a negligible change of pressure ratio. Thus the point of maximum density ratio
must lie at a mass flow less than that of maximum efficiency but greater than that
of maximum pressure ratio.
Now consider a compressor delivering from atmosphere to a large plenum, as
in figure 17.1. If the system is operating near the point of maximum efficiency,
where the density ratio curve has a negative gradient, and the throttle is closed
slightly, the flow through the throttle will reduce immediately, but because the
pressure in the plenum is momentarily constant the compressor will continue to
deliver the original flow to the plenum. This causes an increase of the resident
mass in the plenum, which leads to an increase of pressure and density in the
plenum. The increased pressure will present a greater resistance to the incoming
flow which will therefore diminish. As the pressure in the plenum increases it
will cause the outlet flow through the throttle to increase. The compressor
delivery flow and the throttle outlet flow converge with time as indicated in
figure 17.5. There must therefore be an increased resident mass in the constant
volume plenum at time 2, compared to time 1, i.e. the density must be greater.
Because the density ratio v mass flow curve has a negative gradient this can be
satisfied. The hatched area of figure 17.5 represents the increased resident mass
in the plenum.
Now consider the system operating at the mass flow corresponding to
maximum density ratio. If the throttle is again closed slightly the processes will
be the same as previously except that now the hatched area of the right hand
diagram on figure 17.5, representing an increased density in the reservoir, is not
consistent with the reduced density ratio the compressor delivers at the reduced
flow. This suggests that stability will not be achieved at this operating point.
While this argument has not been supported mathematically, it has shown good
agreement with experimental surge measurements as indicated by Miller and
Wasdell (1987). It has the merit that it can be easily applied to predicted
performance data and tends to be somewhat more conservative than most other
surge predictions. Due to the smaller gradients of the pressure and temperature
ratio characteristics as speed is reduced the points of maximum density and
pressure ratio tend to converge at lower speeds. This is also consistent with
Surge 129

experimental evidence which more often tends to indicate a maximum pressure


ratio being achieved at surge for speeds below about 80% of design.

More complex approaches

Whereas the preceding argument is based on continuity alone, more complex


theories to predict surge such as Elder and Gill (1984) have taken account of
fluid momentum, energy, and continuity and show considerable success in
comparisons with experimental results. One drawback, however, is a
requirement to know the individual stage performance characteristics. If these
cannot be predicted with a high degree of accuracy, then experimental
characteristics must be used. This rather begs the question, since if these have
been measured, the experimental surge points are also likely to have been
measured. Other theoretical investigations by Emmons et al (1955), Howell
(1964), and Stenning (1973) have considered the system to consist of an
actuator disc and a duct to represent the compressor plus a plenum. These
analyses are based on considerations of energy, momentum, and continuity, and
tend to indicate the maximum pressure ratio as the limit of stability.

Surge margin

It is essential for the compressor designer to provide an adequate margin between


the normal operating point, or line, on the performance characteristics and the
surge line. This is to allow for transient variations of the surge line, such as may
occur due to inlet flow maldistribution, or temporary changes in tip clearances
due to transient thermal effects. Transient movements of the working line due to
rapid acceleration of a gas turbine also require adequate surge margin, and for
plant compressors similar requirements arise during start up and shut down. In
the longer term, service erosion of the blading may cause surge line deterioration
and accompanying loss of efficiency may raise the working line. All of these
factors require to be assessed in setting the margin required, particularly if there
are any exceptional circumstances such as ingestion of corrosive substances or
solid particles which will cause severe erosion of the blade profiles. Helicopter
engines, for example, are subject to salt water spray in marine operations and
heavy dust and sand ingestion in some other environments.
A definition of surge margin is illustrated in figure 17.6. The margin is defined
by the difference in pressure ratio between the operating point and the surge line
at the same inlet mass flow function, divided by the pressure ratio at the
operating point, which can be written as:

SM = (Rs - R*)/R
130 Axialflaw fans and compressors

At first sight it may be thought that the margin should be defined by the
working and surge pressure ratios at a constant speed, however the most
physically significant parameter is the flow function at compressor exit,
MVT2/P2, because it is this flow function which is controlled by a downstream
throttle. For pressure ratios greater than 3, the surge line has a roughly constant
value of this flow function, and therefore much the same result is obtained at a
constant flow, or along a constant speed line in terms of the change of outlet
flow function. It is more convenient therefore to define the surge margin in terms
of pressure ratios at a constant value of the inlet flow function.
If a polytropic efficiency of 85.7% is assumed and y is taken as 1.4 we can
write:

mVTj/Pj= (MVTj/PMPj/P,)*6

Writing Q2 for MVT2/P2 it follows that*

QisfQiw = (Rw/Rs)5*

Figure 17.6 Surge margin definition at constant inlet flow

For a constant value of the surge margin Rw/Rs is constant and so the
definition gives, to a close approximation, a constant ratio of surge to working
flow function based on the compressor outlet conditions and its value is in the
range from 0.86 to 0.83 for surge margins of 20% to 25% respectively. The
figure of 20% is an approximate minimum for a gas turbine compressor and a
generally acceptable figure would be 25% based on the above definition and
applied at the design speed. It is fortunate that in many cases the maximum
efficiency lies at about this distance from the surge line. Low stagger blading,
that is designs having relatively high values of the work and flow coefficients,
tend to have the maximum efficiency closer to surge and this is a disadvantage as
Surge 131

the operating point has to be positioned where the efficiency is a point or two
lower than the maximum in order to provide an adequate surge margin.
At one time some designers adopted a practice of computing the design at a
higher pressure ratio than the intended operating pressure ratio at the same mass
flow. For example a design to operate at a pressure ratio of 5 might be computed
at a pressure ratio between 5.5 and 6 . This arose from some early designs giving
little or no margin above the design pressure ratio. Now that better data is
available on the effects of aspect ratio and tip clearance this practice should no
longer be necessary. It is certainly undesirable, as it effectively raises the
matching pressure ratio and leads to greater mismatching problems at part speed.
18 Performance presentation

Introduction

It is normal to present the overall performance of a compressor in the form of


pressure ratio and efficiency for a series of constant speeds, plotted to a base of a
mass flow function. The pressure ratio and efficiency are truly non-dimensional,
but the mass flow is generally presented in a quasi non-dimensional form
appropriate to the particular compressor under consideration. One form is to omit
the area term from MVT/AP and use MVT/P. The temperature and pressure used
are the total values at inlet to the compressor. Another form is to use the
corrected mass flow, which is the mass flow in kg/s when the pressure and
temperature at inlet are corrected to standard values, usually 101.3kPa and 288K.
Me* can alternatively be written as MV(T/288)/(P/101.3). T/288 is written as 6

MT1,/t/P1

Figure 18.1 Performance plot with contours of constant efficiency


Performance presentation 133

and P/101.3 as 5 so that the parameter is written as m Vq/5. The form MVT/P is
preferred because it helps to remind the user of the nature of the parameter and
avoids the tendency to overlook the necessity to correct for the inlet conditions.

Performance graphs

Efficiency v pressure ratio

Efficiencies are most often plotted against the mass flow function as in figure
18.2. At high speeds the range of mass flow at a constant speed may become so
small that it is not practical to plot efficiency in this way. A subsidiary plot of
Working Ri

Figure 18.2 Efficiency characteristics

iMfitropJc •fDctoncy %

Figure 18J Efficiency v pressure ratio


134 Axialflaw fans and compressors

efficiency against pressure ratio is often adopted to overcome this, as in figure


18.3 and this is conveniently plotted to the right of a.diagram such as figure 18.1.
An alternative which can be usefully adopted is to plot contours of constant
efficiency on the pressure ratio v mass flow diagram as in fig. 18.1. This is not
suitable where it is desired to compare two sets of characteristics, which is often
the case.

Non-dimensional speed

The constant speed curves are non-dimensionalised by N/VTj where Ti is again


the inlet temperature. Similarly to the mass flow function this can be expressed
as No, or N/V0 where 0 is Tj/288

Surge and working lines

The surge points of each constant speed curve are usually joined with a curve
known as the surge line. In order to illustrate the surge margin and the operating
efficiency a steady state working line can be superimposed on the pressure ratio
characteristics and this can also be drawn on the efficiency curves as in figure
18.2.

Outlet flow function

For some purposes, such as matching of the LP. and H.P. components of a two
spool gas turbine, it is helpful to plot the characteristics to a base of outlet mass
flow function rather than inlet flow function. The constant speed curves remain

N/T,'*
674
648
606

542

0.4 0.5 0.6 0.7 0.8 0.9


Outlet flow function MTa1/a/ P2

Figure 18.4 Pressure ratio v outlet flow function


Performance presentation 135

as N/>/Tt. In this form the pressure ratio curves take on a very different shape,
and the surge line becomes approximately vertical in the higher speed range. The
characteristics of figure 18.1 are replotted to a base of outlet flow function on
fig. 18.4.

Temperature rise ratio

Although not always presented in overall performance plots it can be useful to


plot the temperature rise function AT/Tj, as shown for the same compressor as in
figure 18.1 in figure 18.5. A point to note about these curves is that for each
constant speed there is a minimum value of AT/Ti below which it is impossible
to operate, and this value increases with the speed. This is caused by choking of
the flow in the final stage. As is clear from fig. 18.4 the outlet flow function
increases rapidly as pressure ratio falls and a choking value of MVT/AP will be
reached, usually in the final rotor or stator. When this occurs, further opening of
the outlet throttle will have no influence on conditions upstream of the choke
plane, and further reduction of pressure ratio is due only to increasing loss
downstream of the choke plane The temperature rise ratio and the inlet mass
flow function are therefore constant below the outlet choking pressure ratio. The
pressure ratio from inlet to the choke plane is also constant; in effect the whole
of the blading up to the choke plane is at a fixed operating condition independent
of the outlet throttle setting when this is a larger area than gives the choke
condition.
This effect can be demonstrated if a multistage compressor is mounted on its
own bearings without coupling to any other component and a high pressure air
supply is connected to the inlet with the outlet open to atmosphere. It will be
found that no matter how high the inlet pressure is raised the compressor will not
rotate at more than 10% to 20% of its design speed. This is because without
coupling to any other unit the compressor can neither absorb work in the normal

0.5 1 1.5 2 2.5 3


MTt^/P,

Figure 18.5 Temperature rise characteristics


136 Axialflow fans and compressors

manner or produce shaft woik as a turbine, apart from the friction of its own
bearings. Thus the temperature rise must be zero across the blading and because
of the outlet choking condition this is impossible except at relatively low speeds.
In contrast to this, if a simple jet engine is windmilled in flight the speed can
rise to about 50% of the maximum. In this case the compressor produces a
pressure and temperature rise and the turbine uses the pressure ratio to give a
temperature drop to drive the compressor, even with no combustion taking place.
The inefficiencies of the process are compensated by the ram pressure ratio
produced by the aircraft's flight speed.

Isentropic and polytropic efficiency

The overall efficiency of a compressor is defined as the ratio of the ideal to the
actual work input. Except in those cases where intercooling is employed, the
ideal work input is the work required by an isentropic process between the same
pressures as the actual process. The enthalpy v entropy diagram is used to
illustrate the ideal and actual processes as in figure 18.4 where ABC is the actual
path of a compression from Pi through P2 to P3. The overall increase of enthalpy
is AH and the ideal process ADE has the smaller enthalpy rise AH|.
The enthalpy rise is given by:

A H = JC p d T

The specific heat at constant pressure, Cp, is a variable with temperature, rising
about ten percent through the normal range of temperatures encountered in the
compression of air, as shown in figure 18.5. A mean value for the temperature
range of any particular process will be assumed for present purposes, hence:

AH = CpmAT

and A H j = C pmA T i

where AT* is the temperature rise of the ideal process. The isentropic efficiency
can be written as:

t1u = (Te -T a)/(T c -T a)

where A, C and E refer to the points on figure 18.6. For an isentropic process the
pressure and temperature ratios are related by the expression:

T j/T , = m

Hence T 2 - T , = T , ( O V T , ) - 1 ) = T , t ( P ^ P , ) (T' - 1 )
Performance presentation 137

Therefore tli. = T,((Pj/P 1)<',' ,v' - 1 )/(T j-T ,)

Writing R for P2/P 1 and AT for T2 - Tj gives:

ilu = (R<lrW-l)/(A T /T l)

Figure 18.6 Actual and ideal compression processes

Temperature K

Figure 18.7 Specific heat of air at constant pressure

This is the most commonly used expression for the overall efficiency of a
compressor. It is often loosely referred to as the adiabatic efficiency, but is more
correctly termed the reversible adiabatic or, preferably, the isentropic efficiency.
It should be noted that the pressure v specific volume diagram has not been
referred to in the above discussion. For flow processes, as distinct from the
quasi-static processes of a reciprocating machine, the area enclosed on the
pressure v volume diagram can only be evaluated to give the work done for a
reversible, i.e. ideal, process and not for an irreversible process. On figure 18.8
the full lines represent isentropic processes and the broken lines irreversible
138 Axialflow fans and compressors

Figure 18.8 Process diagrams

adiabatic processes. For the irreversible compression process from point 1 to 3


the area A13B is obviously greater than for the area A12B of the isentropic
process from 1 to 2 and this would appear to be consistent with the
corresponding enthalpy rises of figure 18.8b. However, for expansion the
irreversible process from 4 to 6 indicates a diagram work area B46A which is
greater than the isentropic work area B45A. This is clearly contrary to the
smaller enthalpy drop indicated for the irreversible process in figure 18.8b. The
area enclosed on the pressure volume diagram is therefore only appropriate for
reversible flow processes. (See Shepherd (1956) pl05)
It is also to be noted that the pressures and temperatures throughout are total
rather than static values. If the Mach number of the flow at inlet and outlet of a
compressor were equal and therefore ratios Pj/pj and P2/P2 were equal, then the
pressure ratio and efficiency would be the same whether total or static values
were used. Where the flow has negligible velocity, as in a large plenum, there is
no difference between total and static values. Some industrial installations have
such plenums at inlet and outlet and the performance is assessed by pressures
and temperatures measured in them. The work input to the compressor is a
function of the rise of total temperature and this is invariably used in the
efficiency derivation from temperature measurements.

Polytropic efficiency

Now consider the isentropic efficiency of an infinitesimal pressure rise.

HdT/T = {(P + dPyP)11'’1*^- 1

Hence dT/T «= {(7 - l)/r)Y)dP/P

On integration this gives:


Performance presentation 139

loge (T2/T1) = {(Y*l)/n7}l°g«R

Hence TI = {(y-1 YYllogeR/logeCiyTi)

This form of the efficiency is known as the small stage or polytropic


efficiency. It provides a more meaningful figure for the comparison of
compressors of different pressure ratios. For a constant polytropic efficiency the
isentropic value diminishes as pressure ratio increases. For example if polytropic
efficiency is constant at 90%, the isentropic value is 87.9% at a pressure of 4;
86.7% at 8 ; and 86.1% at 12. So far as the standard of aerodynamic performance
is concerned the constant value of the polytropic efficiency is more meaningful
than the varying value of the isentropic. In gas turbine cycle calculations the
isentropic is more commonly used.
An understanding of the difference between the two forms of efficiency can be
gained with the aid of figure 18.6* where P|, P2, and P3 are lines of constant
pressure. It is important to note that the vertical intercept between these lines
increases as entropy increases. ABC is the locus of actual compression and ADE
is the locus for isentropic compression. The overall isentropic efficiency is given
by AHi/AH
Considering the processes from Pi to P2 and P2 to P3 separately and calling
these stage 1 and stage 2 respectively we have:

For stage 1 r|,i = Ahu/Ah,

For stage 2 r |l2 = Ahi2/Ah2

If T|« = r|., = t1s2 then:

Ah| = AhnAlsi and Ah2 = Ah&Al*!

Also AH = Ah, + Ah2 = (1/ti, XAhu + Ah*)

and so the overall isentropic efficiency can be written:

Tic = AHy(l/n.)(Ahu+Ahi2)

and ricM, = AH/CAhn + Al^)

Since AH| < (Ahn + Ahi2) because of the divergence of the constant pressure
lines towards the right of the diagram it must be that: r\c < rj, The small stage
efficiency, r\n is more commonly called the polytropic efficiency and will be
denoted rip.
A relationship between tjp and x\u can be derived as follows:
140 Axialflow fans and compressors

For an isentropic process:

l y r . = r *»-

and for a polytropic process:

T2/T, = R(T' u'<nrt

The isentropic efficiency is:


Tlc = (Ti2-T,)/(T2-T,)

Figure 18.9 gives a graphical presentation of the above relationship. Note that
the pressure ratio scale is logarithmic in this diagram and the ratio of specific
heats has been taken as constant at 1.4. The use of a mean value of y instead of a
constant makes only a very small difference. For example at a pressure ratio of
20 and a polytropic efficiency of 85% the isentropic efficiency is 77.9% for y =
1.4 and 78.1% for y = 1.384 corresponding to the mean value over the
temperature range assuming 288 K as the inlet temperature. In figure 18.9 the
broken lines are for y = 1.37 and are only shown at higher pressure ratios where
this would be an appropriate mean value.

90

y ■ 1.4 ’ ** rip - 90%


Y - 1.37

’ *T)p - 85%

T|p ■ 80%
70
2 5 10 20 50 100
Pressure ratio

Figure 18.9 Polytropic and isentropic efficiencies

Efficiency assessment

When assessing the efficiency of a compressor from test measurements there are
a number of factors to be taken into account other than the purely
Performance presentation 141

thermodynamic relation of the pressures and temperatures of the compression


process. In most cases the efficiency of greatest interest is the one which relates
the shaft power or torque requirement to the pressure ratio and the mass flow
delivered to the following part of the system. This requires that the torque
needed to overcome bearing friction and possibly disc windage must be
accounted, as also must the work done on any air extracted before the
compressor delivery.
If the shaft torque is measured directly an effective temperature rise can be
computed:
Torque, T= 60W/2tiN

Power, W = MCpAT

AT = 27cNx/60MCp

where N is rotational speed in rpm, Cp is the mean value of specific heat in the
temperature range concerned, and M is the mass flow rate at the compressor
delivery. This effective temperature rise can be compared directly with the ideal
temperature rise required for the measured pressure ratio to give the isentropic
efficiency based on shaft torque.

AT, = T,(R(T' 1Vr- l )

Tl = ATj / AT

Where no mass flow is extracted before the compressor delivery this value
should be close to the value obtained using the measured temperature rise of the
air across the compressor. The small difference will be accounted for mainly by
the bearing friction, since the heat generated by this will mosdy be absorbed by
the lubricant and will not increase the measured temperature rise of the air flow.
When rolling element bearings are employed the bearing friction is likely to be
less than half a percent of the total shaft power.
Disc friction occurs almost entirely on the front of the first stage disc and the
rear of the last stage disc. Conditions between intermediate discs are likely to be
close to solid body rotation and therefore involve negligible friction between the
air and the discs. This means that, as a percentage of the total shaft power, the
disc friction is likely to be greatest for a single stage fan, and to progressively
diminish as the number of stages and the pressure ratio increases. Whether the
heat generated by disc friction is transmitted to the main airflow will depend on
sealing and venting arrangements of the particular machine. Where torque
measurements are not made a mechanical efficiency can be introduced to allow
for the bearing and windage powers. This is likely to be in the region of 98% to
99%, and the shaft power required will be :
142 Axialflaw fans and compressors

W = MCpATAl.

where T|m= mechanical efficiency and AT = measured temperature rise.


For low speed research compressors, such as that employed at Cranfield
University and described by Robinson (1985), it is essential to derive the
efficiency from torque measurement because the temperature rise is only of the
order of 1°C per stage. An assessment of the bearing losses and disc friction is
desirable in order that the true aerodynamic efficiency of the blading can be
assessed. Torque measurements can be made with the blades removed from the
machine and the blade fixing aiTangements smoothed over. The friction torque
absorbed by the dram surface in such a test should be debited to the blading
performance and this can be calculated from the data of Bilgen and Boulos
(1973) and deducted from the total friction torque measured before coirecting
the efficiency. The drum surface friction is likely to be less than 20% of the total
friction torque.

Efficiency with interstage bleed flows

Two quite different definitions of efficiency may be of interest when there is


extraction of air at intermediate stages through the compressor. The isentropic
shaft efficiency as defined above is of primary interest so far as the overall
performance and the input power are concerned. Where the shaft torque is
measured no detailed knowledge of the bleed flows is necessary, provided the
delivery mass flow is known. This involves measurement of the bleed mass
flows if the delivery flow is not measured directly at the compressor discharge.
The isentropic shaft efficiency can be obtained as described above. Where shaft
torque measurements are not made, it is necessary to make measurements of both
the mass flow and discharge temperature of each bleed flow. The total shaft
work required must then be calculated as:

W = MCpATo.+ mj CpATb|+ m2CpATb2+ etc.

where:
M = delivery mass flow
ATo. = overall temperature rise
mi, m2 etc = individual bleed mass flows
ATbi, ATb2 etc. = temperature rise of bleed flows
Cp = mean specific heat for the appropriate temperature range of each flow.

The effective shaft temperature rise is given by:

ATeff = W /M Cp
Performance presentation 143

and so: T]s T|(R(t‘ ,Vy- l)/ATgff

While the above derivation of the efficiency is the most commonly required, it
is sometimes of interest to know the efficiency of the compression of the
delivery mass flow. This is given by:

T) = T|(R<T' ,)/r- 1)/ATm

Since the objective of low speed interstage bleeds is to improve the surge
performance and this is achieved by bleeding to increase the flow through the
early stages and thus unstall them, the aerodynamic efficiency of their
performance is improved. When the efficiency of compression of the delivery
flow has been restored to values close to those achieved at higher speeds, where
the stages are well matched, the amount of bleed is likely to be near optimum.
In plotting the performance characteristics where there are one or more
interstage bleed flows it is best to use a flow function based on the delivery mass
flow but the inlet pressure and temperature as the base, i.e. M2VTj/ P| rather than
the inlet flow. For a gas turbine this gives approximately the same operating line
irrespective of the bleed flow, whereas if the inlet flow is used the position of the
operating line on the characteristics is dependent on the quantity of the bleed
flow.
19 Stage matching and
surge control

Introduction

The stages of a compressor are only correctly matched at one point on the
performance map. The further away from this point the compressor is operated
the greater is the mismatching of the stages. At one extreme one or more stages
will stall and as discussed in the previous chapter the compressor surges; at the
other extreme choking will occur in the first or last stage. These limitations may
preclude satisfactory operation of the compressor requiring special arrangements
such as variable stagger stators, interstage bleed or twin shaft arrangements to be
provided.

The stage matching problem

A typical map of overall compressor performance is shown on figure 19.1. The


maximum overall polytropic efficiency occurs at the point marked ‘3’, which in
this case lies at a speed of approximately 93% of the design value. It is
reasonable to assume that all the stages must be close to their individual
maximum efficiency conditions on their stage characteristics at this overall
operating point. The individual stage characteristics for the first, middle, and last
stages are also shown in figure 19.1. These are plotted in terms of the pressure
rise coefficient against the flow coefficient. The maximum efficiency points are
marked ‘3’ to correspond with the overall operating point ‘3’. A curve can be
drawn through point ‘3’ on the overall performance map for which the density
ratio across the compressor is constant. The ratio of the inlet to outlet axial
velocities is also constant on this line. For continuity of mass flow we can write:

M = piAiVai —p2A2Va2
Stage matching and surge control 145

where M is the mass flow rate, p is the air density, A is the annulus area of the
compressor and Va is the axial velocity. Subscript i denotes the inlet plane and 2
the outlet plane.

Hence: Vaj/Va2 = (p^piXA^A,)

Using the gas law: p/p = GT and noting that A^Aj is a constant for a given
compressor we have:

Vajt/Vai = (P2/PiXTi/T2)*constant

The polytropic relationship gives:

P2/P 1 = OVTi)* where p = T|py/(y- 1) = 3 (approx.)

Thus Vai/Va2 = (P2/P|)2/3 * a constant

Figure 19.1 Stage matching on overall and stage characteristics

Strictly, the pressures and temperatures in the above expressions should be the
static values; for present purposes it is sufficient to assume that the ratios of total
values are equal to the ratios of static quantities. The polytropic efficiency has
been assumed constant in the above expressions and if this were the case the line
of constant velocity ratio would be at a constant pressure ratio. Since *3* is the
point of overall maximum efficiency the efficiency must fall to either side of *3*
on the line of constant density ratio. As indicated the line curves upwards to the
left and right of ‘3’ when the true efficiencies are accounted for. The line
146 Axialflow fans and compressors

through the point *3' defines points of the same velocity ratio as the maximum
stage efficiency points. All stages therefore have their flow coefficients changed
in the same proportion at any point along this line. Where the line intersects the
surge line at point *5* all the stages would be at their peak pressure rise points
simultaneously, assuming they all had the same ratio of stall to maximum
efficiency flow coefficient Points *1* to ‘6 * correspond to points of diminishing
flow coefficient on the first stage characteristic. Point ‘1’ lies on the 100%
design speed characteristic at the same density ratio as point ‘3’. The flow
coefficients of all stages are therefore increased in the same ratio and all stage
operating points are similar amounts to the right of the maximum efficiency
points.
Point *2’, also at 100% of design speed and on the turbine operating line, has a
greater density and pressure ratio than point *1*. The flow coefficient therefore
diminishes more on each succeeding stage. The mid stage is at peak efficiency
and the last stage is to the left of peak efficiency. Point *4* is at surge at design
speed. All stages are now to the left of peak efficiency and the last stage is at its
stall flow coefficient As discussed previously all stages are at their stall points
simultaneously at point *5’.
Point *6 *lies on the operating line at 70% of design speed and the density ratio
is much reduced from point *3’. The flow coefficients therefore rise rapidly
through the stages and the first stage is well into the stall region while the last
stage is approaching its choking flow coefficient.
If the annulus areas were progressively reduced through the stages to give a 5%
reduction at outlet the matching pressure ratio would be raised to that of point
*2'. This could provide a small increment of efficiency at point *2*, but at low
speeds the mismatching of the stages would be more severe, i.e. the velocity
ratio would be further from the optimum for a given density or pressure ratio.
The maximum efficiency for a given percentage of design speed would be
reduced and the low speed surge line would tend to intersect the operating line.
This is the main reason why designers choose to place the front stage towards its
choking flow coefficient at the design point and the rear stage toward its stall
point. These movements of the stage operating points relative to each other
become greater, for a percentage speed change, the higher the design pressure
ratio. However, the surge line above the matching density ratio (point *5*) tends
to be dictated by the stall point of the rear stage. Below the matching density
ratio the surge line corresponds approximately with the middle stage operating at
its stall point, which necessitates the stages ahead of it operating in the stall
regime.
The stage matching can be illustrated in another way by plots of the relative
flow coefficient, <)>/<J>m, as in figure 19.2 against stage number for various
operating points, where <|>= Va/U, and <J>ra is its value at maximum efficiency.
The choke line indicates the relative flow coefficient at which the stage
characteristic becomes vertical and no further increase is possible. At low
density ratio it is the choking of the rear stage which limits the flow and causes
the front stages to operate in the stalled regime. For compressors having a design
Stage matching and surge control 147

pressure ratio less than 5:1 these mismatching effects are tolerable, but at higher
design pressure ratios they cause such severe stalling of the early stages at low
speeds that arrangements have to be made to alleviate the adverse effects on
blade vibration and surge line shape The surge line problem can manifest itself in
a ldnk as indicated in figure 19.3. This is due to the onset of stall in the first
stage, or to nearly simultaneous stall of the first few stages. In severe cases it can
prevent acceleration of a gas turbine as the steady state equilibrium running line
may intersect the surge line, as indicated in figure 19.3. Since transient
acceleration requires operation above this line, surge free acceleration becomes
impossible.

Figure 19.2 Relative flow coefficient through the stages

Figure 193 Surge line kink

Bleed off-take

Two different methods have been adopted to alleviate the vibration and surge
problems of part speed operation of a single shaft compressor. The first and
simpler, but less efficient, is to bleed air from the compressor, either at an
intermediate stage or at the exit before the combustor of a gas turbine. The effect
of exit bleed is to lower the operating line at a given mass flow. It is
148 Axialflow fans and compressors

thermodynamically inefficient because all the work done in compressing the


bleed flow is generally wasted, although in some cases a proportion of the
energy is recovered, as when the bleed flow is exhausted to a by-pass duct and
some thrust is produced.
Intermediate stage bleed has the effect of (installing the early stages due to the
increased flow induced through them. This improves their aerodynamic
efficiency. The thermodynamic inefficiency is less than for exit bleed since the
bleed air is only compressed through about half the number of stages. The net
result on a gas turbine cycle is therefore much better than for exit bleed.
In deciding at what stage to apply bleed to a compressor a number of
considerations arise. First the bleed port must be large enough to pass the
required flow. The size required for a given percentage of the flow will diminish
the further towards the rear the port is placed due to the increasing air density.
From an efficiency point of view the further forward the port the better. In
general it has been expedient for these reasons to place an intermediate bleed
port near the mid stage. In some cases two or more stages have been provided
with bleed ports

c Reduced loss bleed off-take


Figure 19.4 Bleed off-take arrangements

A number of designs of bleed port have been used and some of these are
indicated diagrammatically in figure 19.4. All appear to work satisfactorily and
no preference can be given from a performance point of view but designers may
have a preference from a structural viewpoint. In a case where a large quantity of
bleed flow was required for aircraft services the arrangement of figure 19.4c was
adopted in order to reduce the bleed flow pressure loss. There was concern that
the compressor performance would be adversely affected when bleed flow was
not in use, but this proved unfounded. A somewhat similar bleed port design is
described by Tubbs and Rae (1991).
Stage matching and surge control 149

Variable stagger stator blades

Figure 19.5 Vector geometry with variable stagger stators

The second method of alleviating low speed stall problems is to fit variable
stagger stator blades to one or more stages. The effect of such variables on the
vector geometry is indicated in figure 19.5.
It is apparent that at a given axial velocity and blade speed the inlet angle
relative to the rotor blade can be restored to near the design value by increasing
the IGV stagger. If the following stator is set to an increased stagger its inlet
angle can also be near optimum. By progressively reducing the increment of
stagger angle from stage to stage for about half the total number of stages it is
possible to have all early stage blade rows operating at near optimum inlet angles
at speeds well below design. As speed is reduced the increments of stagger angle
must be increased, as indicated by the variable stagger schedules of figure 19.6.
Since all blade rows operate near their optimum angles the efficiencies achieved
at pan speeds are close to those achieved at design speed. The stage pressure rise
and flow coefficients fall markedly as stator stagger angles are increased, as
illustrated by the stage characteristics for various stator stagger angles of figure
19.7. This has the effect of reducing the mass flow and pressure ratio at a speed.
For a jet engine this results in a much smaller range of rotational speed being
required for a given increment of thrust. The engine therefore responds more
rapidly to throttle movements as less rotational inertia has to be overcome to
achieve the same thrust increment This effect is greatest at high altitude where,
150 Axialflow fans and compressors

due to the reduced air density, the gas torques of the compressors and turbines
are reduced but the rotational inertia is not A very kinked surge line can be
smoothed by multi-variable stators to the extent that no surge problem is
encountered on acceleration. In some gas turbines the use of variable stators and
bleed have been combined, for example by using only a variable inlet guide vane
plus an interstage bleed valve. The Rolls Royce Avon was an early example of
this form of off design control.

Increase

SUzzer
angle " r
increase

% Design speed
Figure 19.6 Typical variable stator Figure 19.7 Stage characteristics for
schedule variable stagger stators

Industrial axial compressors

These are often driven by constant speed motors. To obtain a range of mass flow
at constant pressure ratio it is a common practice to fit variable stators to this
type of compressor. Typical performance characteristics are shown on figure
19.8. It will be noted that the surge line for the constant speed variable stator
performance is much less steep than that for the fixed stator variable speed
characteristics. This can be most readily understood by considering a single stage
characteristic in terms of the pressure rise and flow coefficients. For fixed
geometry variable speed the stall points for a range of speeds will fall on a
parabolic curve because Ap/pU2 and Va/U are constant for the stall point and
therefore Ap = Ki U2 and Va = K2U and so Ap = KjVa2 where K|, K2, and K3 are
all constants. For variable stator constant speed operation it can be assumed that
the stall points of the different characteristics at various stator settings will all
occur at approximately the same rotor inlet air angle, and also rotor outlet angle.

Hence: tan ai - tan a 2 = AH/UVa = a constant

At constant speed it therefore follows that:


Stage matching and surge control 151

Figure 19.8 Performance at constant speed with variable stagger stators

AH = KVa. and: AP = K^pVa

Hence, if the efficiency variation is small, the stall points will lie
approximately on a straight line through the origin. Matching calculations for a
fixed stagger rotor and a variable stagger stator show a tendency for the two
blades to mismatch as stator stagger is varied but this is quite a small effect for
the first 20° of stator movement and so the above argument is a good
approximation.
Variable stator blades are usually fitted with a circular platform which rotates
in a bore in the casing. The outer end of the boss is fitted with a lever which is
connected to a ring encircling the casing. Rings for several stages may be ganged
together and a common ram operates all the variable stator rows simultaneously.
The lever geometry of each stage is so designed as to give it the appropriate
angle change relative to its neighbours. The operating ram is connected to a
control system which determines the angle of the vanes, usually as a function of
N/VT, for a gas turbine, although pressure ratio has been used in some cases.

Twin spool gas turbines


152 Axialflow fans and compressors

An alternative arrangement for alleviating the matching problem of high pressure


ratio axial compressors in gas turbines is to use two compressors in series on
separate shafts, each driven by its own turbine. In aero engines the shafts are
concentric and the high pressure turbine drives the high pressure compressor,
while the low pressure turbine drives the low pressure compressor by the inner
shaft The arrangement is illustrated in the diagram of figure 19.9.
This arrangement not only allows some alleviation of the low speed surge
problem, but also offers a considerable increase of the design pressure ratio
available from a given number of stages. Consider a single shaft compressor
having ten stages each having a temperature rise of 45°C when operating at the
design speed with an inlet temperature of 288K. The rise would be 450 °C and
making the approximation that R - OVTO3 gives:

R*{1 +450/288 }3= 16.8

If, say, four stages are on the LP shaft and six on an independent HP shaft a
higher pressure ratio can be achieved.

For the LP shaft: Ru» - {1 + 180/288 }3 = 4.29

The stages on the HP shaft are raised in blade speed and appropriately modified
in annulus dimensions so that they are matched to theLP outlet flow with the
first HP stage running at the same Mach number as the firstLP stage.This
means a higher stage temperature rise:

AT/per stage = 45(1 + 180/288) = 73.125°C

Hence: AT for 6 stages = 438.8

The HP inlet temperature = 180 + 288 = 468K

Hence Rn> = (1 + 438.8/468 }3 = 7.27

Hence the overall pressure ratio is:

Rc = 4.29 *7.27 = 31.2

The HP pressure ratio is the same as would be obtained from six LP stages
because the same blade sections are running at the same Mach numbers whereas
when all are on one shaft the rear six stages only produce a pressure ratio of
3.92. As illustrated by figure 19.10 the maximum pressure ratio is obtained when
the stages are equally split between LP and HP shafts. Thus at the expense of the
complexity of two concentric shafts and the associated bearing arrangements a
Stage matching and surge control 153

No. of LP stages

Figure 19.10 Effect of LP/HP shaft split on overall


pressure ratio often stages

significantly higher pressure ratio may be obtained from the same number of
stages, or the same pressure ratio may be obtained from fewer stages. It is
emphasised that this is only because the HP stages are raised in Mach number so
that the first HP stage is running at a similar Mach number to the first LP stage.

Off-design operation of twin spool gas turbines

The off-design operation of twin spool compressors can be illustrated by an


approximate analysis which assumes that both turbines operate with choked
nozzles and the downstream nozzle is also choked, plus the assumption that the
turbine and compressor efficiencies are constant along the steady state operating
lines. While the assumption of choked nozzles only becomes inappropriate at
low pressure ratios, the assumption of constant efficiencies may appear too
sweeping, however, the inaccuracy is small enough to justify the approximation
for demonstration purposes.
The condition for choked nozzles gives:

Q* = Qj * Q6

where Q = MVT/AP and the subscripts refer to the planes indicated on figure
19.10. The mass flow can be taken as equal at all three planes, and the area of
each nozzle is fixed, and so:

(T4/T5),/2(Pj/p4) = A4/A5 = a constant

The temperature and pressure ratios are connected by:


154 Axialflaw fans and compressors

Hence for constant values of y and T| the turbine temperature and pressure ratios
will be constant so long as the nozzles are choked.

T4/T3 = a constant; and P4/P5 = a constant

Similarly Ts/T6 and P5/P6 are constants. From the work balance between the HP
compressor and its driving turbine we have:

T3- T 2 = C,(T4-T5)

and similarly for the LP system:

T2 - T, = C2(T5 - T6)

where the constants C| and C2 account for the differences in specific heat
between compressor and turbine gases and the net change of mass flow due to
the addition of fuel and the extraction of cooling flows. Hence we can write:

M(T3-T2)l/2/P3 = MVT4/P4[C,aVT!)/T4]l'2P4/P3

P4/P3 is the pressure ratio across the combustion chamber, with a value in the
region of 0.95 and is approximately constant

Hence: M(T3-T2 )m!?z is approximately constant

MViyP2 = [M(T3-T2)‘'2/P3] (P3/P2)(T 2/(T3-T2) ) '/2

Writing Rh = P3/P2 and putting tiy/(y*1) = 3, gives:


Rh = (T3/T2)3

Hence: MVTyP2 = Constant * Rh/(Rh10- 1)w


To obtain the LP pressure ratio corresponding to any HP pressure ratio we
write:

(T3-T2)/(T2-T1) = [ C ,( T 4 - T 5 ) ] / [ C 2 ( T 5 - T 6)] = [ ( T 4 - T O /T 4 ] [TV(T3-T6)] (T4/ t 5)

= a constant
Hence [(T3-T2)/T2] [Ti/(T2-Ti>] OVTi) = a constant
(T3 -T 2)/r 2 = RH10- 1 and (T2 • T,)/T, = Rl10 -1

Hence Rl i,3(Rh1'3- I VCRl1"- 1) = a constant


Stage matching and surge control 155

And so: RL= 1 /(1 - K(RH,/S-1)}3


The constants can be determined by inserting the design values for the flow
functions and pressure ratios. These relationships allow the steady state
operating lines of the HP and LP compressors to be plotted on their respective
performance characteristics. As an example, assume both HP and LP
compressors have a design pressure ratio of 4.0, and their inlet mass flow
functions are expressed as a percentage of their respective design values. The
result is shown on figure 19.11 where it can be seen that while the HP operating
line falls steeply as pressure ratio and mass flow are reduced, the LP line is much
flatter and tends to converge on the surge line at low mass flows, even although
the surge margin at design speed has been chosen to be considerably larger than
is usual. However, surge free operation might just be possible, which it most
certainly would not be, for the same overall design pressure ratio of 16, if a
single shaft compressor without variable geometry stators or bleed were adopted.
In practice the HP compressor would be designed with a considerably lower
surge margin at design speed, and a variable stagger IGV might be used on the IP
compressor, or a small percentage of airflow might be bled from the inter
compressor duct at low speeds.

% Design inlet flow function

Figure 19.11 HP and LP operating lines for a two shaft jet engine
Other applications
It is worth noting that the above analysis applies also to the IP and HP
compressors of a three shaft turbofan, but not to a two shaft turbofan where the
LP turbine drives the fan as well as the LP compressor. The analysis for the HP
flow and pressure ratio also applies to a simple single shaft jet engine
156 Axialflow fans and compressors

compressor, although it must be remembered that it is only satisfactory when the


final propulsion nozzle is choked.

Active surge control

In recent years experiments have been made into means of actively controlling
the onset of surge by sensing the initiation of the rotating stall which, it appears,
precedes the surge phenomenon itself. This has generally been done by
application of pressure transducers close to the stage where the rotating stall
commences. Day (1991) used the signals generated to operate 12 air injection
valves spaced around the casing to produce damping of the stall to successfully
delay surge in a low speed research compressor. A number of other research
projects have reported similar results, while Cargill and Freeman (1990), who
have conducted surge investigations on a simple jet engine at high speed, warn
that results of research on low speed rigs may not be directly applicable to
engine conditions. They indicate that to be successful the damping control will
require to operate on the initial rotating stall disturbances which grow into surge.
To do this an extensive number of sensors and actuators may be required.
Further work may find practical ways to achieve this, and on a small engine with
a centrifugal compressor Ffowcs-Williams, Harper and Allwright (1992) report a
10% increase of power being obtained before surge.
20 Inlet flow maldistribution

Introduction

Axial compressors have proved to be sensitive to certain types of inlet flow non­
uniformity. The first serious problems due to this arose in some early military jet
aircraft due to firing rocket propelled missiles. Incompletely burned gases from
the rockets were ingested by the engine and caused it to surge. This was
diagnosed to be due in part to non uniform pressure and temperature at the
compressor inlet and in part to the incompletely burned gases acting as
additional fuel in the combustor. The latter effect raised the compressor working
pressure ratio, while the former lowered the surge pressure ratio, resulting in an
engine surge. An expedient to avoid the problem was to link the missile firing
trigger to the fuel flow control so that the engine fuel flow was momentarily
reduced as the missiles were fired. Other aircraft of the same vintage suffered
surge due to flow separation from the intake walls when static or at very low
forward speed on the ground, particularly in a cross wind. Most of these aircraft
had wing root intakes on a swept leading edge which were more prone to
separation than the intakes of more modem pod type installations.
Research testing on the problems were initially concentrated on distortion of
the total pressure at the compressor inlet face. It soon became apparent that when
the total pressure varied radially, but was uniform circumferentially, the
deterioration in performance, and surge line in particular, was much less than
when the pressure varied circumferentially but was radially constant. Real flows,
of course, varied in both directions, but the separation into radial and
circumferential variation was a useful simplification both for test purposes and to
aid understanding. Because it was the more damaging, circumferential variation
received greater attention and a basic understanding came with the concept of the
compressors in parallel theory. While this concept is a considerable
simplification it does provide a useful approximation to reality and for this
reason is widely accepted.
158 Axialflow fans and compressors

Compressors in parallel

The first published reference to this basic understanding of circumferential inlet


distortion was given by Pearson and McKenzie (1959). It is postulated that
circumferential mixing or attenuation of the flow distortion cannot take place
within the bladed annulus due to the relatively large number of blades in each
row and the small axial gaps between the rows. A number of identical
compressors are assumed to operate in parallel, and deliver to a common outlet
plenum. It is therefore assumed that all the compressors deliver to the same static
pressure, but have different inlet total pressures. The compressors are assumed to
have identical performance characteristics in terms of pressure ratio, (this is
expressed as outlet static/inlet total pressure), efficiency, and mass flow per unit
inlet annulus area. They may be of various sizes to represent sectors of the real
compressor having an approximately constant inlet pressure. The sum of their
inlet areas is equal to the inlet area of the real compressor. With these
assumptions the average pressure ratio and total mass flow can be calculated for
a known inlet pressure pattern, and a known performance in uniform inlet flow.

J*L

Mass flow-*

Figure 20.1 Compressors in parallel simulation of inlet maldistribution

The diagrams of figure 20.1 illustrate the situation for the simplest case where
one half of the annulus has one uniform pressure and the other half a different
uniform pressure. The more usual total / total pressure ratio can be obtained by
adding the average dynamic pressure at outlet to the static outlet pressure.
Inlet flow maldistribution 159

That these assumptions are reasonably accurate has been demonstrated by a


test of a moderate pressure ratio, nine stage compressor with a 90° sector throttle
plate fitted at the inlet. The inter-stage static pressures at the casing were
measured at a number of circumferential positions as shown on figure 20.2.
Points to note are:

• The inlet static pressure varies almost as much as the total pressure,
indicating that the inlet velocity is almost uniform.
• The difference between the maximum and minimum static pressures
circumferentially increases from the compressor inlet to stage 6 inlet and
reduces rapidly through stages 8 and 9 to become approximately constant at
outlet
• The circumferential extent of the distortion increases from 90° at inlet to
about 160° at stage 9 entry.

4r

0.5 — 1— •— 1---- ■— --- 1-----*----- •— 1----■---- — >


0 90 180 270 360
Circumferential station - dagraaa
Figure 202 Circumferential variation of stage static pressures
with 90®sector inlet spoiler

The first point can be explained by considering a compressor operating on a


vertical part of a pressure ratio v mass flow characteristic with inlet distortion of
the total pressure. If two parallel compressors are considered, they must both
operate with the same mass flow function, since this is the only value possible at
this particular speed. The unique mass flow function corresponds to a unique
velocity since the total temperature is constant.
The second point can be demonstrated by considering the static pressures
through the stages of the compressor when operating with uniform inlet flow at
two pressure ratios corresponding to the operating points with a distorted flow.
In figure 20.3 the maximum and minimum pressures at each stage from figure
160 Axialflow fans and compressors

Out

3 3.5 4 4.5 5
Outlet static / Intel total p rw u f

Figure 203 Design speed stage static pressures with uniform inlet flow

20.2 arc represented by lines 'A' and ’B\ which correspond to the overall pressure
ratios with maximum and minimum inlet total pressures respectively.
In figure 20.4 the difference between the maximum and minimum stage
pressures derived from figure 20.3 are compared to the values given by test from
figure 20.2. The good agreement suggests that the assumptions of the
compressors in parallel theory give a good approximation to the flow behaviour.
The third point indicates that the effective proportion of the annulus area
occupied by the low pressure inlet flow increases through the stages. This is
inconsistent with the theory, but would tend to cause a greater pressure ratio for
the low pressure inlet flow as the axial velocity would fall more rapidly, giving
more pressure rise from the rear stages. The corresponding reduction of annulus
area occupied by the high pressure inlet flow would reduce its pressure ratio.

Stag* inlet

Figure 20.4 Comparison of test and deduced pressure differences

The test performance with uniform flow and distorted flow are compared with
the performance predicted by the theory in figure 20.5. The prediction is
Inlet flow maldistribution 161

% Design mass flow


Figure 20.5 Test and predicted performance with distorted inlet flow

pessimistic and does not give the increased range of mass flow at a constant
speed observed on test. Despite these deficiencies the effect on the surge line is
at least a first approximation to the experimental results.

Distortion parameter

More extensive research on the subject was reported by Reid (1969). In


particular Reid examined the effect of varying the extent of the flow distortion,
and derived the parameter DQo which has proved valuable to define the severity
of the flow pattern, however complex it may be. The parameter is determined by
the 60° sector of the compressor inlet annulus which has the lowest average total
pressure, which is denoted by Pm. The average total pressure of the complete
annulus is PAv, and the mean inlet dynamic pressure is P av - Pav. Since the static
pressure may vary over the annulus the dynamic pressure is calculated from a
knowledge of the total mass flow and the average total pressure, PAv. Thus we
can write:

Dc« = (P av - P«o)/(Pav - P av)

This has proved a practical means of defining the severity of an inlet distortion
and has been used as a basis for contractual agreement between aeroengine and
airframe manufacturers. Care has to be taken if intake suction tests are used to
determine a value for Dc«) as these would normally have a uniform static
pressure across the inlet/engine interface plane, whereas the compressor or fan
can induce a variable static pressure as demonstrated in figure 20.2. This means
that the static pressure gradients along the surface of the intake may differ in the
162 Axialflow fans and compressors

two cases and in consequence boundary layer separation may not be the same.
The resulting DQo in the engine may therefore be different from that measured
on an intake suction rig test. The only totally satisfactory method is to conduct
the intake tests with a representative compressor in position when measuring the
intake/engine interface conditions. Unfortunately this is not always practical at
an early stage in the development programme.
The value of DQo which can be tolerated for an acceptable loss of surge
pressure ratio tends to increase as the overall pressure ratio increases. This is
obvious from the consideration that in the compressors in parallel theory the
deficit in pressure has to be restored within the blading. Thus a given deficit of
pressure may be impossible to restore in a single stage fan, but could be a minor
matter for a multi-stage compressor with a design speed pressure ratio of 10. It is
also important that there should be no point within the blading where the static
pressure has free circumferential communication, other than inter-blade row gaps
which typically do not exceed 25% of blade chords. For example, if an inter­
stage bleed off-take has a manifold close to the annulus and no circumferential
partitions, the static pressure may become circumferentially uniform and the
pressure deficit has to be restored in only half the total number of stages. This
would cause a more serious loss of surge pressure ratio, which could be reduced
by placing partitions at a number of points around the manifold so as to prevent
free communication of the static pressure.

Inlet temperature distortion

The compressors in parallel theory can readily be applied to the case of


temperature variation over the inlet face of the compressor. For the case where
the total pressure is uniform it indicates that equal pressure ratios have to be

Common rpm

= 288K
N/VT, = 100%

; I Average performance
with temperature
maldistribution

Inlet mass flow —►


Figure 20.6 Inlet temperature maldistribution
Inlet flaw maldistribution 163

generated by the various sectors of the compressor which are operating at


different values of N/VT( as indicated in figure 20.6. Obviously the sector
operating with the highest Ti will have the lowest surge pressure ratio, and will
be the critical one. If inlet pressure and temperature both vary this can also be
assessed by applying the varying pressure ratios at the appropriate non-
dimensional speeds.
It will be apparent that the theory will indicate a variation of outlet temperature
even when the inlet temperature is uniform. For a two shaft compressor system
this implies a variation of inlet temperature to the second compressor. Whether
this will diminish significantly due to mixing will depend on the design of the
connecting duct. It is common for such ducts to be relatively short in comparison
to their circumference, and also to have between four and eight struts which will
prevent complete circumferential mixing.
21 Compressor rig testing

Introduction

It is a feature of the gas turbine that the major components can be tested
separately. This has major advantages for the compressor as it is inconvenient to
assess the full range of the performance characteristics in an engine, e.g. the
margin of pressure ratio from the equilibrium running line to the surge line is
critical to satisfactory transient operation but difficult to assess at all speeds in an
engine. Any rig test, however, is only an approximation to the performance in
the engine, albeit often a very close one.
The high levels of power required to test compressors at sea level inlet pressure
are often not available, and inlet throttling is required to reduce power and
therefore the Reynolds number is also reduced. Inlet total pressure may not be
uniform over the compressor inlet face on the engine due to intake flow
separation caused by cross wind effects during ground running, or aircraft
manoeuvres during flight. Intake simulators may have to be used on rig tests to
reproduce the appropriate conditions. Blade tip clearances may differ between
rig and engine due to differences of temperature and rotational speed. All these
and other factors must be accounted for and corrections applied to rig test data
before the best possible representation of performance in the engine is obtained.
Despite these problems rig testing provides a very useful development tool, and
is used by most organisations involved in axial compressor manufacture.

Power requirements

The power required to drive an axial compressor is given by:

W = MCpAT
Compressor rig testing 165

where W is the shaft power, M is the mass flow rate, Cp is the specific heat of air
at constant pressure, and AT is the temperature rise across the compressor. For a
mass flow of lOOkg/s and a temperature rise of 100°C this gives a power of
approximately 10,000kW. Test facilities of this size and greater exist at the
major manufacturers plants, but it will be appreciated that there is a considerable
incentive to reduce power requirements where possible.
Considering the factors concerned, the mass flow can be reduced in two
possible ways. If the inlet pressure is reduced by a throttle valve upstream of the
compressor the mass flow will be reduced in proportion to the pressure for a
constant value of the flow function MVT/P. Thus if the machine has a design
speed pressure ratio of 5, the inlet pressure may be reduced to say one third of
atmospheric pressure. This will allow pressure ratios from about 3.5 upwards to
be tested at the design speed. Allowance has to be made for outlet ducting
pressure losses so a pressure ratio of 3 would not be available. The throttle
pressure drop would be reduced when testing at lower speeds to allow the lower
pressure ratios to be obtained. The power requirement falls rapidly with speed
reduction, in fact approximately as speed cubed for constant inlet pressure and
temperature.
A more extreme method of reducing the mass flow is to manufacture a scale
model of the compressor. If this is made to 1/2 linear scale the mass flow will be
1/4 of full scale and the power required will also be 1/4 of full scale. Further
power reduction by inlet throttling of the model is possible. Although great care
is necessary to produce scale models which are in all respects aerodynamically
accurate the use of models for very large components such as high bypass ratio
fans has proved satisfactory.
Both in throttling and in scaling down, the major aerodynamic parameter of
concern is the Reynolds number. Unless the test value is close to the operating
value significant corrections will be necessary to account for this effect, as
described in Chapter 10. In aero-engine practice the lower Reynolds numbers
due to reduced pressure at high altitude are often of most concern. In this case a
throttled inlet rig test may give a more appropriate value of Reynolds number
than one at sea level atmospheric inlet pressure.
Another method of reducing power is to run the rig tests at a lower inlet
temperature than appropriate to the normal engine operating value. This occurs
naturally if the high pressure component of a two or three shaft engine is tested
at atmospheric inlet temperature. For example, if the high pressure component of
a three shaft engine has an inlet temperature in the engine at sea level of 500K
and is rig tested at 288K the temperature rise will be reduced in the ratio of the
inlet temperatures. This is because we have the relationship:

AT/T, = (R'HVtir. 1(

where Ti is the inlet temperature, R is the pressure ratio, y is the isentropic


index, and T} is the polytropic efficiency. Thus if the design pressure ratio of the
166 Axialflow fans and compressors

high pressure compressor is 5, and the efficiency 85%, taking y = 1.4 gives
AT/Ti = 0.687. Hence the temperature rise at the design point in the engine
would be 0.687*500 = 343.5°C, but on rig test it would be 0.687*288 = 197.9°C.
However for a given non dimensional speed parameter N/VT] the quasi-
nondimensional mass flow function mVTj/Pj would be constant and therefore
mass flow would rise inversely as VT falls and would reduce in proportion to Pi.
These effects can be more readily appreciated if the expression for power is
written with non dimensional groups for mass flow and temperature:

W = (M'/Ti/P1XCpAT/Ti)Pi/'/T,

Hence: W/(P,VT,) = (MVT,/P,XCp AT/Ti)

The group on the L.H.S. is a quasi-nondimensional and indicates that the power
is proportional to the inlet pressure and to the square root of the inlet
temperature.
Reduced inlet temperature also has the advantage of reducing the required
rotational speed. Similar Mach numbers are achieved in the compressor when
N/VTi is held constant If an engine has an upper limit to its mechanical speed it
may not be possible to simulate altitude performance at sea level temperatures
because the altitude value of N/VTj may require the maximum N to be exceeded.
Supplying a rig test compressor with cooled inlet air is a means of simulating
altitude conditions without overspeeding mechanically. A cold air supply of the
necessary size requires a major engineering facility but is conveniently available
where a compressor rig test facility is allied to an engine altitude test plant

Variation of specific heat

The specific heat at constant pressure ,Cpt is a function of temperature. For air it
rises from a value of approximately lkJ/kgK at 200K to l.lkJ/kgK at 800K as
shown on figure 18.5. The effect of Cp variation is therefore secondary, but must
be taken into account in accurate analysis of performance data. Some research
testing has been conducted using the FREON refrigerant gases. The advantage of
these is that their low sonic velocity allows high Mach numbers to be developed
at low velocities and rotational speeds. Mechanical stresses are thus reduced,
which eases design problems and reduces the risk of mechanical failure. On the
other hand the gas has to be cooled and recirculated, which leads to a
considerable increase of complexity of the test facility.

Power measurement

Compressor efficiency can be derived from pressure and temperature


measurements alone. While there is no alternative to pressure measurements this
is no real problem as these can be made with high accuracy at all levels of
Compressor rig testing 167

pressure. Temperature measurement is more difficult and accuracy of


temperature rise assessment becomes progressively less acceptable as the rise
falls below 100°C corresponding to pressure ratios less than 2.5 for 288K inlet
temperature. For this reason it is desirable to have an alternative means of
deriving the temperature rise. From the basic equation for power it is obvious
that this can be done if the mass flow and power are measured. Since the mass
flow and rotational speed must be measured for their own sake, torque is the
necessary additional measurement On small low speed research rigs this is
commonly done by mounting the driving motor on trunnion bearings and
measuring the torque reaction directly. For larger test facilities this is not
practical, and a torque meter is required. One type of torque meter consists of a
short length of highly stressed shaft to which an accurate straingauge is attached.
The straingauge gives a measure of the torsional deflection of the shaft which is
calibrated against torque. Another form of torque meter consists of a calibrated
torque shaft with a toothed wheel at one end. The other end carries an outer quill
shaft which has another toothed wheel at its free end, adjacent to the first toothed
wheel. Torque in the calibrated shaft causes a phase displacement of the teeth in
the two wheels. This is measured electronically in modem versions of the
instrument but in early types the measurement was made optically. Both these
types of torque meter can give an accuracy of ± XU% of full scale deflection. It is
desirable to arrange the test facility drive train such that the torque shaft can be
easily replaced. A number of torque shafts suitable for a range of maximum
torques should be provided so that the most suitable can be used for each test

Power sources

Three types of power source are conventionally used for compressor rig test
facilities - electric motor, steam turbine, or gas turbine. Electric drive is
commonly used for powers below lOOOkW and has been used up to 4000kW.
Above this size special supply arrangements may be necessary which are
unlikely to be economic for the relatively low utilisation of the facility, which is
unlikely to exceed some hundreds of running hours per year. Steam turbines
provide a very suitable type of drive provided a steam supply is readily available.
If boilers have to be fired specially for the test facility the low utilisation may be
further reduced by the time necessary to raise steam.
A gas turbine is a convenient form of drive in that start up is only a matter of a
few minutes and no large scale associated equipment such as boilers is required.
They are available in a suitable range of sizes. Rolls Royce first installed a gas
turbine powered test facility in 1951 and chose a similar source for a much larger
new plant in 1980.
It is not uncommon for compressor test facilities to be part of a group of
facilities providing test arrangements for combustors, turbines, and engine
altitude testing. These facilities require large air supplies and it can be
convenient to use the compressed air supply through a turbine to drive a
168 Axialflow fans and compressors

compressor test facility. The air supply may be heated in a combustor to increase
the turbine power if desired.

Facility layout

Inlet arrangements

The general arrangement of a typical test facility is shown diagrammatically in


figure 21.1. Atmospheric air is drawn in through filters to a plenum. The filter
banks should be arranged symmetrically to the downstream flow in order to
avoid creating whirling flow in the ducting. A well flared bellmouth takes the
flow smoothly from the plenum to the intake ducting system. Suitable
dimensions for the flare are shown in figure 21.2. These are important as the
mass flow measurement is made a short way downstream of the throat of the
flare. A well proportioned flare will give a thin boundary layer at the
measurement station and a discharge coefficient in excess 0.99. The exact value
of the discharge coefficient can be determined by pitot tube traverses across the
boundary layer.

Mass flow measurement

The mass flow is measured by the mean of a number of static taps on the duct
wall as indicated in figure 21.2. The pressure drop across the filters or the
absolute pressure in the plenum must also be measured. Alternatively a fixed
pitot tube Beading outside the duct boundary layer will give the same result.

Inlet Return bleed — Outlet throttle


throttl / distributor / “f 1! _
J .— __ .__ . /unit /Torquemeter

'Airmeter / r~ \ \ . . box ““
1Air filter Bleed control Smoothing \ Outlet
throttle Metering screens plenum
orifice
Figure 21.1 Test facility arrangement

The size of the airmeter measuring plane should be chosen so that the Mach
number does not exceed 0.4. This results in a maximum static depression of
1500 mm water gauge (110 mm of mercury or 15 kPa). The lower limit of the
Compressor rig testing 169

measuring depression is dependent on the accuracy of the measuring system but


should be such as to maintain an accuracy of at least

Figure 21.2 Airmeter flare dimensions


Adaptedfrom Rolls Royce internal report

Downstream of the airmeter the duct is of increasing diameter to suit the inlet
throttle valve. The length from the flow measuring plane to the throttle must be
sufficient to prevent any upstream influence on the flow in the measuring plane
whatever the throttle setting. A length equal to at least three diameters of the
measuring plane should be provided.

The inlet throttle

The throttle may take a number of forms two of which are illustrated in figure
21.3. Note that these are such as to avoid creating whirl in the downstream duct
or to deflect the flow to one side of the duct as would a butterfly type throttle.
The throttle would normally be driven by electric motor, except for very small
facilities where it may be hand operated.

The inlet plenum

It is necessary to allow the flow to settle downstream of the throttle, and for this
purpose the ducting is further enlarged downstream to give a mass average
velocity in the region of 15m/s. A number of baffle plates are fitted in the large
diameter pipe and possibly a honeycomb straightener if there is a possibility of
whirl being present At design speed the axial velocity at entry to the compressor
blading will be in the range 100 to 200m/s so there will be a large acceleration
from the inlet pipe to the blading. This will help to maintain a uniform total
pressure across the inlet face of the compressor for normal testing. If testing with
non-uniform inlet total pressure is required provision should be made to fit
distortion producing screens some way ahead of the blading but after most of the
acceleration has taken place. If placed where the velocity is too low the screens
170 Axialflow fans and compressors

will not be effective. Inlet total pressure and temperature must be measured
ahead of the blading. In the absence of distortion screens the inlet ducting
arrangements described should provide a sufficiently uniform total pressure such
that only 3 or 4 total pressure probes are necessary. With distortion screens an
extensive array of probes is required. Alternatively an area traverse may be

Jen \

\ /

- r r r r r t :

4 * i± i, *:

Multi vane type

Figure 21.3 Two types of inlet throttle


made, but this is inconvenient and time consuming for routine testing.

Inlet temperature

Total temperature is normally constant from atmosphere to the compressor inlet


face. Where this is the case thermometers are often placed in the plenum behind
the filters, which has the advantage of being a region of very low velocity. In
some test arrangements it is necessary to return some compressor bleed flows to
the inlet pipe. These add heat to the inlet flow, therefore the compressor inlet
temperature must be measured downstream of the bleed flow re-injection point.
It is important that the bleed flow should be completely mixed with the main
inlet flow before the temperature measuring plane, and to ensure this a
distributor pipe as indicated in figure 21.4 can be used. Apart from measurement
problems, a non-uniform inlet temperature will adversely influence the
compressor performance similarly to a non-uniform total pressure.
Compressor rig testing 171

Outlet measurements

At the compressor outlet the essential measurements are again total pressure and
temperature. The total pressure generally has significant radial variation, usually
being highest near the blade mid-height. The total temperature on the other hand
tends to be highest near the blade ends. Radial rakes of 5 or 6 limbs are therefore
desirable to obtain a reasonably representative average value for both pressure
and temperature. About 5 rakes spaced circumferentially around the annulus
insure against any lack of axisymmetric flow and allow a good overall average
from the 25 or 30 readings for each parameter. Area averages of the
measurements are commonly used but mass or momentum averages are preferred
by some engineers. The differences are generally small but can be significant for
single stage fans where the dynamic pressures are a high proportion of the
overall pressure rise.
An alternative and very simple means of obtaining the outlet total pressure is to
have a short duct of constant annulus dimensions behind the blading. This should
preferably be at least two annulus heights in length with five static taps
circumferentially spaced on each of the annulus walls half way along its length.
The total pressure is calculated from the average of the static pressures as
described in Chapter 11.

Discharge arrangements

Downstream of the outlet measuring plane it may be desirable to reproduce the


flow path of the intended installation e.g. the diffuser and combustion chamber
head in the case of a gas turbine. This is desirable in so far as these features may
influence the flow upstream in the final blade rows of the compressor. Beyond
this it is usual to discharge the flow into an outlet plenum and from there through
one or more throttle valves to atmosphere. A silencing muffler may be fitted to
the final discharge pipe, if required for environmental reasons. The positioning
of the outlet throttles relative to the compressor can influence the frequency of
the surge cycle and the ease of recovery from surge. If the throttles are fitted at
the exit of a large outlet plenum the surge cycle frequency will be low. As the
plenum volume is reduced the frequency increases and at some limiting plenum
size the compressor may appear to stabilise in a deep stall condition. This can be
a dangerous condition where large amounts of energy are absorbed by
recirculation of air within the blade rows of the compressor, leading to large and
rapid increases of temperature which have been known to melt even steel
blading.

Test procedure

It is convenient to conduct the testing at predetermined values of the non-


dimensional speed parameter NWT|. Having brought the compressor close to the
172 Axialflow fans and compressors

required speed, it is necessary to measure the inlet temperature and calculate the
exact rpm for the chosen value of N/VlV It is usual to carry out the acceleration
with the throttles set to operate well clear of surge. A preliminary calculation of
the required inlet pressure versus speed to keep within the power available
should be made, and the inlet throttles adjusted to follow this pressure/speed
relationship as the plant is slowly accelerated. The permissible inlet pressure
should make allowance for the increase of power towards surge, since it is
generally more convenient to conduct each constant speed curve at a constant
inlet pressure.
If bleed valves or variable stagger stators are fitted it is necessary to specify a
schedule for their settings against speed. For a new design on initial test these
may have to be decided on past experience of similar compressors, or possibly
on the output of performance predictions where these are available. For
subsequent development tests the analysis of the initial performance will provide
further guidance for the bleed valve and blade settings. The outlet throttles are
set simultaneously during the acceleration to follow a pressure ratio v speed
relationship estimated to be well clear of surge. If necessary this pressure ratio v
speed relationship should also be set so as to be greater than may encounter any
choked flutter vibration of the blading.
In running each constant speed curve it is best to start at the lowest pressure
ratio required and take a full set of measurements. The outlet throttles are then
closed to take the pressure ratio to surge. Again a previous estimate of the
probable surge conditions should be made. The outlet throttles should be closed
more slowly as the expected surge point is approached. Surge may be detected
audibly, but microphones may be necessary in the test cell if the control room is
remote or well sound proofed. A sensitive test for surge is the oscillation of a
manometer or pressure gauge attached to an outlet pressure tapping or to an
airmeter pressure tapping. A snap measurement of both of these pressures is
desirable to locate the surge pressure ratio and mass flow. At high speeds the
pressure ratio measurement is often the more important as the mass flow
variation may be small. At low speeds the mass flow measurement may be the
more important as the curve of pressure ratio against mass flow often tends to be
relatively flat towards surge.
Having located the surge point, suitable intervals of pressure ratio or mass flow
can be determined for making full sets of measurements, including a set as close
as possible to the surge point. The outlet throttles will have been opened as
rapidly as possible after locating the surge point, and it is convenient to
commence with the lowest pressure ratio required. At least six points are
desirable to define the characteristic fully. It is necessary to allow time at each
point for temperatures to stabilise.
There is no particular order in which the different speed curves must be run,
but it may be convenient to run them in either ascending or descending order. If
there are known or suspected regions where dangerous blade vibration may
occur these may be left until last. On the other hand, if strain gauges are fitted to
some blades, it may be best to carry out a survey for areas of high stress initially,
Compressor rig testing 173

as the gauges may be prone to failure after prolonged running, particularly if the
compressor is surged frequently. Regions of high stress may then be avoided for
extensive running during the full performance calibration.
The number of speed curves required to fully define the performance will vary
depending on the particular requirements of the test, and the application of the
compressor. Typically curves may be run at intervals of 10% of the design speed
from 30% to 80% and then 5% intervals to 110% if this is mechanically safe.

Measurements and analysis

Manual recording of manometers, thermocouples and other gauges has given


way over the last several decades to automatic data recording using scani-valves,
transducers etc. linked to a digital computer which is programmed to record and
analyse the data to give any desired output. This has introduced the possibility of
recording and analysing much more data than was practical previously. For
example with manual recording it took a skilled team of five or six testers about
one hour to run one speed curve. Using a slide rule and tables or graphs it took
one engineer about the same time to analyse the measured data and plot the
overall performance curves for one speed. The analysis of interstage static
pressures to obtain individual stage characteristics took much longer, typically
two or three man weeks depending on the number of stages and the number of
test speeds analysed, with the result that these characteristics were not analysed
for all tests. With automatic data recording and computerised analysis this
situation has been revolutionised, and where it is justified the analysed
performance can be available in real time. The same is true for detailed traverse
measurements where the traverse probe is computer controlled to move from one
position to another, recording and plotting the measurements of total and static
pressure and airflow angle automatically.
Further information on compressor testing is to be found in Dimmock (1961)
and the ASME power test codes for compressors and exhausters.
22 Stage performance prediction

Introduction

Many methods are available for the prediction of stage performance on a two
dimensional basis. They all depend on two basic requirements; the first is a
knowledge of the variation of the air outlet angle as a function of the inlet angle,
and the second is the variation of the losses or efficiency, again as a function of
the inlet angle. The first requirement determines the work characteristic, while
the second gives the pressure rise. The method described by Howell (1945) is
typical of the application of correlated cascade data to produce stage
performance predictions. By introducing secondary losses and a work factor, the
data was corrected to give an approximation to three dimensional stage
performance, based on the mid-span blade geometry for moderate to high
diameter ratios. For low diameter ratios (less than, say, 0.6) Howell’s method can
be used for the various radial blade sections as described in Chapter 15.

A preferred method

Another method, also originally described by Howell and Bonham (1951), has
been modified by the author to produce a very simple and direct method for the
prediction of stage performance in terms of work, pressure rise and flow
coefficients..
The optimum (maximum efficiency) air angles are determined for rotor and
stator by the methods of McKenzie (1980) as summarised in Chapter 4. The
outlet angle of the preceding stator is also determined. The vector diagram
quantities for the maximum efficiency conditions are thus known and the work
and flow coefficients can be calculated. If the optimum angles for rotor and
stator are not consistent with a common flow coefficient an average value should
be used. If they are far from consistency it indicates some radical mismatch in
the design, and the performance prediction is unlikely to be valid. A difference
Stage performance prediction 175

of 5% in the optimum flow coefficients is the maximum to be tolerated. The


rotor and stator optimum flow coefficients are given by the relationships:

For the rotor Va/U = l/(tanao + tanaO

and for the stator VaAJ = l/{ tan a 2 + tan a 3}

where cto is the outlet angle of the preceding stator, <X| is the inlet angle to the
rotor, &2 is the outlet from the rotor, and a 3 is the inlet angle to the stator.

The work characteristic

Howell and Bonham gave data for the slope of the work characteristic at the
design or maximum efficiency flow in terms of the relative work and flow
coefficients, y/\|fd and <p/<p«). For the case of constant outlet angles it is readily
shown that the slope is given by:

Tan0 = (l/y d) - 1

where \|fa is the maximum efficiency value of the work coefficient. When the
slopes of the work characteristics of some low speed compressor tests were
examined in these terms it appeared that the slope was reduced as a function of
increasing space/chord ratio. An approximate relationship for the slopes was
deduced as:

Tan © = {(lA|/d)-1}{1 - 0.4S/C}

Arguably the S/C should be the average of the rotor and preceding stator
values, but the rotor appears to predominate, and where rotor and stator values
differ a value biased towards the rotor S/C should be used, say:

S/C = 12(S/C)r + (S/C)s) / 3

Although the equation for tan@ is linear, it gives the work characteristic with
an acceptable accuracy for low space/chord ratios, as will be shown below.
A.R.Howell's curve for deflection variation with incidence can be drawn as in
figure 22.1 to show the variation of outlet angle. Since the angle variation is
given as a ratio to the nominal deflection, high deflection blades will give a
larger variation of outlet angle from nominal incidence to maximum deflection
than a lower nominal deflection blade design. The latter is likely to have a
greater space/chord ratio than the former and intuitively might be expected to
have the greater variation of outlet angle.
The curves presented in figure 22.2 give an alternative means of determining
the variation of the outlet angles. These have been derived from the off design
176 Axialflow fans and compressors

(i - i*)/e*
Figure 22.1 Outlet angle variation derived from Howell’s e/e* data
Adaptedfrom Howell (1945) with permission of I Mech E

-0.6 -0.4 -0.2 0 0.2 0.4 0.6


tan a, - tan at*
Figure 22J2 Outlet angle variation derived from low speed
compressor tests

performance of the compressor tests reported in McKenzie (1980), and show the
space/chord ratio as being the major parameter determining the rate of change of
the outlet angle. Figures 22.3a and 22.3b show work coefficient characteristics
for typical nominal air angles calculated using four methods:

1) Constant outlet angle


2) Tan 0 = {(1A|/d -1} {1 -0.4(S/C)}
3) (a2- a 2*)/e*] = /{ (i- i*)/e*]); where e = <*i - a 2 (See figure 22.1)
4) tan a 2 - tan a 2* = /(tan ai - tan ot|*) (See figure 22.2)

Figure 22.3a shows the results for typical nominal angles at a S/C of 0.5 and
figure 22.3b for the same nominal angles at S/C = 1.0. The constant outlet angle
assumption obviously gives the same result at both values of the space/chord
Stage performance prediction 177

ratio, as also does the Howell data (method 3) since it is independent of the
space/choid ratio.
While methods 2 and 4 are almost the same as each other at S/C = 0.5, they are
distinctly different at S/C = 1.0. Howell's data gives a lower value of the work
coefficient towards stall than the other methods at S/C = 0.5, but higher than
method 4 at S/C = 1.0. These two methods would be similar from nominal to
stall at S/C = 0.75. From nominal towards choke Howell's data is exceptional in
that it indicates a minimum value of the outlet angle when the value of (i - i*)/e*

0.4 0.5 0.6 0.7 0.8 0.9 1 0.4 0.5 0.6 0.7 0.8 0.9
Va/U Va/U

Figure 213 Work v flow coefficient for various outlet angle variations

= -0.25. However, with the exception of method 2 the others are very close to
the line of constant outlet angle for S/C = 0.5 at flow coefficients above the
nominal, but this is not so at S/C = 1.0. It would be surprising if the S/C ratio did
not have a significant influence on the variation of the oudet angle, and since
both methods 2 and 4 were derived from compressor tests, rather than cascade
tests, they are perhaps more likely to be realistic. Method 2 is therefore
suggested for approximate work, particularly when the space/chord ratio is less
than 0.75. Method 4 is recommended as the most accurate for a wide range of
space/chord ratios. Checks at other nominal air angles give similar variations
between the methods.

Efficiency and pressure rise

Having determined the work coefficient it is necessary to know the total pressure
losses or the efficiency of the blading in order to calculate the pressure rise
characteristic. If the efficiency characteristic can be determined directly then the
pressure rise function is immediately available as the product of efficiency and
work coefficient. Where the total pressure losses are known, as say from cascade
tests, the process is less direct and is exemplified by the method described by
A.R.Howell (1945). Howell gives a unique curve for the profile drag coefficient
178 Axialflow fans and compressors

in terms of the function (i- i*)/e*. The annulus friction drag and the secondary
drag, given by Cd.= 0.02S/h where h is the blade height, and Co, = 0.018Q.2
where Cl is taken as the theoretical lift coefficient given by:

Cl = 2(S/CXtan ai - tan ai)cos On

which ignores the drag term Cd tan Om. The total drag is given by:

Cd= Cdp+ Cd* + Cd,

The loss coefficient can be found from:

Cd = (S/CXAPl/0.5p V!2)cos3 a„, /cos2ai

Writing GJ = AP/0.5 p V,2 and the ideal static pressure rise coefficient as:

Cpi = 1 - (cos <Xj / cos Cb)2

the efficiency can be written as:

T| = 1 - tD/Cpi

This value of efficiency can be used as the stage efficiency for a 50% reaction
stage, but for other reactions the values for rotor and stator should be calculated
separately and weighted to account for the reaction. This can be inconsistent, for
example at 100% reaction since Cpi is zero the efficiency for the stator is minus
infinity. Alternatively the actual static pressure rise of each blade row can be
calculated and the sum taken as the stage total pressure rise, which is only
strictly correct if the axial velocity is constant across the stage. Another, and
perhaps preferable, alternative is to note that the total pressure losses may be
subtracted from the stage work to give the stage total pressure rise:

Stage AP = pAH - APlr - APls

Note that the density p must be introduced to give the pressure terms the same
dimensions as AH. Care must be taken to convert the denominators of the loss
coefficients from relative inlet velocities to blade speeds to give the expression
in the form:

Stage AP/pU2 = AH/U2 - AP^ /pU2 - AP^ /pU2


Stage performance prediction 179

Other methods

Leiblien's correlation

Many other systems and correlations have been developed since Howell's data
was published, and as is to be expected the degree of sophistication has tended to
increase. Space will only permit mention of a few of the systems which have
been developed for the prediction of loss or efficiency.
Leiblien (1956) developed a loss parameter which was correlated as a function
of the diffusion factor for the reference minimum loss incidence.

Leiblien loss parameter = 0.5QJi(S/C)cos a 2(cos a^cos aO2

Its value rises gradually from about 0.005 at Dp = 0.1 to 0.02 at Dp = 0.6. Above
this value of the diffusion factor the loss parameter rises rapidly and Dp = 0.6 is
taken as a design limit, while 0.45 is considered a conservative design value.

Wright and Miller

Wright and Miller (1991) present a correlation for the profile loss and separately
an end wall loss parameter. The profile loss is correlated in terms of the Lieblein
loss parameter as discussed above but related to the equivalent diffusion ratio
(Deq) and the effect of inlet Mach number. D«, was a development of the earlier
diffusion factor taking account of the blade thickness/chord ratio.

Endwall loss parameter =: €3** (h/c)(V|/V2)2 = /(e/C and loading)

Figure 6.6 illustrated this end wall loss correlation and the influence of aspect
ratio and tip clearance

McKenzie

A method based entirely on low speed compressor tests was presented in


McKenzie (1980) and the presentation was improved in McKenzie (1988).
Figure 4.3, taken from the latter shows contours of constant efficiency plotted on
a vertical axis of Cp, and a horizontal axis of S/C. The spine of the contours,
giving the maximum value 91% efficiency, is given by the relationship:

S/Cop, = 9(0.567 - CPi)

It must be noted that, by the design rules of the same paper, when the air outlet
angle is less than 20° it is not practical to design blading at the optimum S/C as
the camber becomes excessive and the incidence too large a negative value. Thus
a lower than optimum S/C has to be adopted which moves the design to the left
180 Axialflow fans and compressors

on figure 4.3 with a consequent loss of maximum efficiency. Being obtained


directly from compressor tests these efficiencies include all loss sources.
However they require to be adjusted for changes of Reynolds number, aspect
ratio, dp clearance and Mach number. The test value of Reynolds number was
1M05 the aspect rado was 2, the tip clearance 1.8% of blade height, and the
Mach number about 0.2. The data of Chapter 10 can be applied to correct for
Reynolds number differences, and Chapter 6 for aspect ratio and tip clearance
corrections. The decrease of efficiency with increase of Mach number can be
estimated with the aid of the data of Chapter 9.

Off design efficiency

The off-design efficiency may be estimated from a curve presented by Howell


and Bonham (1951) which is reproduced as figure 12.2. The base for the graph is
the value of AH/UVa relative to its value at the maximum efficiency point. This
curve was derived from low speed research compressors. A slightly modified
version derived from other sources including high speed multi-stage compressors
has been added. This gives lower efficiencies particularly at the low work and
high flow end of the characteristic. The efficiency is given as the ratio of the
local value to the maximum, the latter being determined from one of the methods
discussed above. The total pressure rise coefficient is given by the product of
efficiency and work coefficient
23 Overall performance
prediction

Introduction

The prediction of overall compressor performance with an acceptable accuracy is


highly desirable in order to avoid costly and time consuming development if the
initial performance of a new design is deficient Many systems have been
developed for this purpose, with varying success. There are two basic systems.
The first is known as stage stacking and, as the name implies, it depends on a
knowledge of the individual stage characteristics and a method of synthesising
them to give the overall performance. The second depends on correlations of
overall performance of existing designs expressed in terms of a number of
fundamental parameters which, given the design geometry of a new compressor,
allow the overall performance to be predicted.

Stage stacking

Where the individual stage characteristics of each stage of a compressor are


known, either by prediction or measurement on a similar compressor, the overall
performance can be calculated by estimating the outlet conditions of the first
stage for chosen conditions of mass flow, speed, inlet pressure and temperature.
The output is then used as input to the second stage and so on through all the
stages. This process is known as stage stacking. The steps in the calculation are
as follows:
Given:
N, M, P,and T,
The annulus dimensions at entry to each stage.
The mean radius flow angle leaving each stator.

MVTi/(A|Cosao P i) gives Vai/(VTiCosao) from compressible flow relations.


182 Axialflow fans and compressors

From the known speed N and the annulus dimensions the mean blade speed U
is calculated. Hence Vai/U is obtained, and from the known stage characteristic
the pressure rise coefficient and the efficiency are found. The pressure rise
coefficient can be written as:

Vp = CpT, (r^ 1* - 1)/U2

where r is the stage pressure ratio, which can now be calculated, and the stage
temperature ratio follows from:

i y r , « i + (i/nXifr,yM )

The outlet flow function from the first stage, which is the inlet flow function to
the second stage, is:

MVTyAjcosctjK = (MVt ,/A,cos otiP,) ( 1/rXTj/Ti )in(A,cosa,)/(Ajcosaj)

Thus the calculation can be repeated for the second stage and all subsequent
stages. The overall pressure ratio is given by:

R = rjr2 ..... rn

The outlet temperature from the final stage allows the overall temperature rise to
be found from:

A T = T (n + 1) - T i

The overall isentropic efficiency can now be calculated from the usual
expression:

nu = (R<1tW- 1)/(AT/T|)

The whole calculation must be repeated a number of times for different mass
flows at the same speed, and then for a range of speeds to give the gamut of
characteristics. The appropriate mass flows at each speed are quite restricted. If
too large a value is chosen the flow coefficient will become excessive after a
number of stages indicating choking. Too small a mass flow at high speeds will
lead to the rear stage flow coefficient falling into the stalled regime. At low
speeds the front stages may be stalled, in this case however the calculation
should continue as the compressor as a whole may not be in surge.
Howell and Calvert (1978) describe the results of a computer code developed
for stage stacking which also develops the required stage characteristics from
annulus and blade geometry. Basically the mid height blade geometry is used,
but correction factors are included for diameter ratio, radial variation of axial
velocity, and scale effects, amongst others. The correlations involved were based
Overall performance prediction 183

largely on high speed compressor test data including transonic fans. The
predicted results shown in their paper for four compressors ranging in design
pressure ratio from 1.88 to 10.26 are in good agreement with the experimental
measurements except for the prediction of too high surge pressure ratios in the
upper end of the speed range for the two higher pressure ratio machines.

Overall performance correlation

An alternative approach is based on the direct correlation of the overall


performance of existing compressors. Howell and Bonham (1951) developed
methods of using the characteristics of an existing compressor to predict the
overall performance of a similar new one. The method was particularly
recommended where the new design was a modification of an existing
compressor with one or more stages added or removed. Robbins and Dugan
(1956) attempted an extension of Howell and Bonham's method using more
generalised correlations. This was not particularly successful, although their
concepts of basing the prediction around the matching point, i.e. the point of
overall maximum polytropic efficiency, and the line through the maximum
efficiency points at all speeds are useful.
A method based on correlations of overall performance which attempts to
utilise some of the best features of Howell and Bonham as well as Robbins and
Dugan, has been developed by the present author. The match point as used by
Robbins and Dugan is the focal point to which the correlating parameters are
related. The work was based on the performance of six compressors of design
pressure ratios between 2.5 and 5.0. It was not thought appropriate to involve
compressors of higher design pressure ratio as these most often involve the use
of variable stators or interstage bleed at low flows and pressure ratios, and this
effectively means a different compressor for each configuration so far as
correlation parameters are concerned.
A primary part of the correlation is plots of AT/N^2 against (Va|Van)1/2/N,ei,
where the suffix ^ indicates division by the parameter's value at the match point
Vai and Van are the mass average axial velocities at inlet to the first rotor and
exit from the last stator respectively. These parameters will be written as and
O mri respectively. <l>Mrei is a form of the flow coefficient based on the geometric
mean of the inlet and outlet axial velocities. A typical example of the resulting
straight line relationship between these parameters is given on figure 23.1. The
slope will be denoted by tanO. Recalling that for the stage work coefficient the
slope at the point of maximum efficiency was given by:

tan© = { (!/¥ * ,)- 1){1 -0.4S/C}


184 Axialflow fans and compressors

It can be shown by considering an operating point at a speed infmitesimally


different to the speed at which the match point occurs* and at the same overall
temperature rise, that:

tan 0 —(AT] tanOi + AT2 tan0j + etc)/ATo,

where the subscripts u , etc indicate the stage numbers. AT] etc. are the stage
temperature rises at the match point, and ATM is the overall temperature rise at
the match point. It is assumed that the efficiency is the same as at the match
point

1.3
N/Nom
0 102%
□ 100%
A 95%
X90%
*89%
• 80%
♦ 75%
♦ 60%
0 54%
0.7

0.6
0.8 0.9 1.1 1.2
$Mr*(
Figure 23.1 Typical plot of and Q rm

The polytropic efficiency at the match point is denoted T|~ and the maximum
efficiency at any speed by T|*. The ratio t \A \~ was plotted against (Vai/Van Xd
where again indicates the ratio to the value at the match point. The
significance of the velocity ratio is that it indicates the degree of departure of the
front and rear stage incidences from their optimum values, bearing in mind that
the optimum incidences occur at 4 ^ and Oma equal to unity and the mid stage
of the compressor is at optimum incidence at all speeds for this condition. This
also implies that the peak efficiencies at all speeds lie on a line of constant
AT/N2. A satisfactory correlation of the maximum efficiencies was obtained in
this form and the resulting curve is shown on figure 23.2
The variation of efficiency with flow and pressure ratio at any speed was
correlated in a similar form to that used by Howell and Bonham, who presented
a curve of ryt|* against x/X ’ where x = AH/(UVa) and * indicates its value at
maximum efficiency. This was for individual stages. For overall multi-stage
performance the equivalent parameters, as introduced above, are used in the
Overall performance prediction 185

(VS|/Vfla)r«|
Figure 23.2 Maximum efficiency Figure 23.3 Efficiency variation at
v velocity ratio constant speed

form WI>Mret As indicated on figure 23.3 the resulting curve is somewhat


narrower than Howell and Bonham’s, particularly at the choke end, i.e. at low
values of'Fja/tlWj

The calculation

It is first necessary to determine the values of (AT/N2)*, {(VaiVao)l/2/N}\


(Vai/Van)*’ and T|* at the match point. If all the blade geometiy and annulus
dimensions are known it is possible to estimate the optimum incidence on to the
first and last rotors at their respective mid-span radii. The mean radius flow
angle leaving the IGV and penultimate stator then allow the optimum Va/U to be
determined for the first and last stages. Allowing for any difference in mean
radius it is simple to calculate the optimum value of Vai/Va„. Since Va„ is the
optimum axial velocity at inlet to the last stage it is necessary to make an
approximate estimate of the axial velocity ratio across the last stage to determine
the overall velocity ratio Vai/VaQThe density ratio at this velocity ratio is then
given by:
Pn/Pi = (Ai/An)(Vai/Van)

The temperature rise

Ignoring the difference between total and static densities, a first approximation to
the total temperature ratio is:
To/r, = (p4>i)1'2

For a known inlet temperature the temperature rise can now be found. At the
match point all stages can be assumed to operate at their optimum efficiencies.
186 Axialflow fans and compressors

The optimum value of AH/U2 is therefore calculated for the mid span of each
stage from the blade geometry. These are expressed as values of AT/N2 and
summed to give the value of overall AT/N2 . Since a value for overall AT has
already been obtained, a value for N* the speed at the match point, can be
derived. For most aero engine type compressors the match point speed lies
between 90% and 95% of the design speed. This results from the tendency to
place the front stages on the choke side of maximum efficiency at the design
point in order to improve the low speed surge line shape, and also from designers
choosing to operate at above critical Mach numbers on early stages. For
industrial compressors the situation could be different and the designer may seek
to design at the match point to maximise the efficiency.
Having determined an approximate value for the match speed and temperature
rise the inlet and outlet velocities are calculable. This allows the axial Mach
number at both inlet and outlet to be determined and the ratio of stagnation
densities across the compressor to be calculated. A second approximation to the
temperature rise can then be obtained, and this also requires a better
approximation to the optimum efficiency. This is obtained by applying the
methods described in Chapter 22 to each stage. Individual stage pressure ratios
are obtained using the stage temperature ratio and the optimum efficiency with
the appropriate value for the ratio of specific heats. The overall pressure ratio
and temperature ratio give the optimum value of efficiency assuming a mean
value for the ratio of specific heats.

Constant speed characteristics

All the necessary values for the match point and the work characteristic gradient
are now known. A number of speeds are selected such as 1.2, 1.1, 1.0, 0.9, 0.8
etc. times the match speed, to as low a value as desired. Intermediate speeds at
1.15, 1.05, etc. may be desirable in the upper speed range. The procedure has
been applied successfully to very low speeds such as 0.2 times the match speed,
which are not always suitable to test without special instrumentation, but are
useful for engine start-up studies.
For each constant speed curve:

a v r ,) ' = i + (a t /n 2x n / Vt ,)2

A first approximation assumes po/pi = (Tq/Ti) which is equivalent to assuming


T| = 86% approximately.

Hence : (Vaj/Van)* = (Tn/Ti)*2(An/AO

(Vai/VanXd = (VaI/Van)’/(VaI/Van)**

= /(Vai/Van)rei from figure 23.2


Overall performance prediction 187

and 71* = tT ( ti7 0


R = (T n /T |)* p where (3 = x\V(Y-1)

(Va/Van) = R ( T |/T n X A n /A |)

(Vai/Van)«i = (Va^VaD)V(Va^VaD)“

Hence another value of tTAi ~ is read from figure 23.2. If this differs from the
previous value by more than 0.5% further iterations should be made until
agreement is obtained. At each speed a series of arbitrary values of <X>Mret are
chosen. Values from 0.7 to 1.2 usually cover the required range, and intervals of
either 0.1 or 0.05 are satisfactory depending on the detail of characteristic
definition required. For each value of $Mni the calculation proceeds as follows:

4'ld = l+ ta n © 0. ( l -<!>,**)

Hence CP/<X>M)iei is evaluated, and iyr|* is obtained from figure 23.3 and so r\ is
derived from:
= oi/n V

AT/N2 = MVKAT/N5)’

Th/Ti = 1 + (AT/N2XN/n/T,)2

r = a v r i ) 'n'(’M)

1st Approximation : Vaj /Van = R(Ti / T nXAn/A|)

(Va,/Va„ )W/N = <t>Mrd{(Va,.VaII)'/3/N |‘

Va,/N = {(Val.Van),n /N)(Va^Va0)l/1

V a^Ti = (Va^N)(N/VT,)

MVTj/A|P| = /(Va^/TD

MVTB/Ab Po = 1(MVTi/A|P, )IK)(TiJT{)mAJAa

Va0 /VTn = /(MVTd/Aq Pn )

2nd. approximation:

Vai/Va„ = {(Va^/T,)/(Va0/VTn))(T1/TB)1'2
188 Axialflow fans and compressors

Vai/VT, = ((VaiVaB)l/I/N)(Vai/Vao)in(N/VT|)

MVTi/A, P, =/(Vai/VT|)

MViyP, = Ai(MVT^A|Pi)

Hu = (RT"l'1r- D/AT/T,

When the values of Vai/Vag for a constant speed curve are examined it is
found that they give a maximum value at a greater mass flow than corresponds to
maximum pressure ratio. This is markedly so at high speeds, but maximum
pressure ratio and maximum velocity ratio tend to converge as speed is reduced.
Maximum velocity ratio corresponds to maximum density ratio and as discussed
in Chapter 17 this is an indication of surge.

High speed corrections

At the upper end of the speed range the predicted constant speed characteristics
may give a maximum mass flow, the flow reducing at lower pressure ratios as
indicated in figure 23.4. This indicates choking in the final stage, and the real
characteristic is taken as having the maximum flow constant as pressure ratio is
further reduced. The temperature rise is also taken as constant as pressure ratio
falls below the value corresponding to maximum flow, (figure 23.5) and
efficiency values can be calculated from this temperature rise and arbitrary
pressure ratios. The explanation for this is that when the rear stage chokes
further reduction of the overall pressure ratio is due only to increasing losses in
the outlet stator, and the operation of the compressor proper, in front of the
choked plane, is unaffected by downstream changes such as opening of the outlet
throttle.
A further correction is necessary to adjust for choking of the front stage at high
speeds. From a knowledge of the first rotor throat width at its mid span the
relative inlet Mach number and inlet flow angle to just choke the rotor passage
can be calculated and hence an estimate of the choking mass flow for each speed
can be made. Where this mass flow is less than given by the procedure of the
preceding paragraph it is accepted as the predicted value. As indicated on figure
23.4 the pressure ratio v mass flow curve is blended into the basic predicted
curve if the latter falls to less than the choking flow before reaching surge.
Otherwise the characteristic is completely vertical and surge is determined by the
maximum velocity ratio Vai/Van in the same way as previously discussed. The
efficiency values on the choked part of the characteristic are obtained from the
assumption that the temperature ratio Tq/Ti will be the same as for the non-inlet
choked characteristic at the same value of the outlet flow function.
Overall performance prediction 189

m V T b /P b = (M /V T i/P ,X P i/P n X T n A 'i)1/l = ( M V T ^ > [)d .(P i/P n )d ,(T 0/ T I) ' /3

Hence: (Pn/Pi)d, = ((MVTyPO/CMVTi/POd,)(Pn/Pi)

It is apparent from this relationship that the line joining corresponding points on
the original and choked pressure ratio characteristics is a straight line through the
origin; Pq/Pi = 0; MVTj/Pi = 0. The efficiency at a choked pressure ratio can now
be calculated from he choked pressure ratio and the unchanged temperature
ratio.

260 90

240 80
c
70
£ 220
Si
cUi
2 60

| 200 50
3 180
40
I
160 30

140 20
2 2.5 3 3.5 4 4.5
Pressure ratio

Figure 23.4 Choke corrections to mass flow and efficiency

Pressure ratio

Figure 23.5 Choke corrections to temperature ratio


190 Axialflow fans and compressors

Test comparison with prediction

The predicted and test characteristics of a six stage compressor are shown on
figures 23.5 and 23.6. This is the same compressor as used to demonstrate the
linear relationship between 4/ibj and <X>Mni in figure 23.1. At the higher speeds it
is clear that the maximum pressure ratio occurs at a significantly lower flow than
maximum density ratio and the latter is the closer prediction of surge.

% Design (low

Figure 23.6 Predicted and test pressure ratio characteristics

70 r > » > i t i > i » i * i— i i »— i


30 40 50 60 70 80 90 100 110
% Design flow

Figure 23.7 Predicted and test efficiencies


24 Performance with altered gas
properties

Introduction

To this point air has been assumed as the working gas, however, it is necessary
to consider the effect of the variation of the properties of air on the design and
operation of compressors, and also, for industrial compressors, the use of gases
other than air.
Air has a fixed gas constant of 287 J/kgK but the specific heat at constant
pressure increases with temperature. When designing at specified conditions the
appropriate values of Cp and y (= Cp/Cv) can be chosen, and varied from stage to
stage through the machine as necessary. However, an aero engine compressor
might be designed for a subsonic aircraft speed at high altitude using an inlet
total temperature of 250K, but could be operated supersonically at low altitude
with an inlet total temperature of 430K. The corresponding value of Cp at the
compressor inlet would increase by a little less than 1.5% in the latter case, and
this would have some small but measurable effect on the performance.
For the high pressure compressor of a three shaft turbofan the mean value of Cp
at altitude cruise conditions might be 1076 J/kgK. At sea level take off this value
would change by less than 0.2%. However, if this compressor is rig tested with
normal room temperature at inlet, then the mean Cp for the temperature range of
the compression would fall to 1012 J/kgK, i.e. a drop of almost 6%, which
should certainly be accounted for in test analysis. Clearly, for an industrial
compressor operating with entirely different gases the change of Cp could be
much greater and there will also be a change of the gas constant
It is obviously desirable to be able to estimate the performance of a compressor
operating with a gas of different properties to the one for which it was designed,
or with which it was tested. This chapter gives an outline of methods by which
approximate corrections can be made and indicates some of the problems
involved in the compression of different gases.
192 Axialflow fans and compressors

The significance of density ratio

Moyes (1956) investigated the case of a change of specific heat of air. He


assumed that the polytropic efficiency of the compressor would be the same
when it was operated so that the incidence to all blade rows was the same with
both values of Cp This requires that the flow coefficient onto each stage remains
the same when the gas properties change. The continuity equation then shows
that the density ratio will be the same with the two sets of properties:

M =piA|Vaj = p2A2Va2

P2/P 1= (A.yA.XVaj/Vas) = Constant * (Va,/U,)/(Va2/U2)

This implies that if a set of performance characteristics are plotted in the form
of density ratio and polytropic efficiency against first stage flow coefficient they
will apply whatever the gas properties. This leaves the question of the equivalent
rotational speeds for the two gases and this requires that Mach numbers should

4 r 100*
Efficiency
80
Ui
60

2 -

0.3 0.4 0.5 0.6


Flow coeff. Va, / U

Figure 24. Performance plot suitable for any gas

also be the same. This is not possible throughout the compressor because the
Mach number will fall at a different rate through the stages due to the different
temperature rises required for the same density ratio. Since the first stage Mach
number is the greatest, and therefore the most critical, the closest approximation
to the performance with another gas will be obtained when the first stage Mach
numbers are the same. This will be achieved by running at the same values of
N/V(tG0 or U/V(*yGt), where U is the first stage tip speed, if the compressor is
scaled. The static temperature is necessary in this parameter for strict equality of
the Mach numbers as the ratio T/t for the same Mach number will differ for
gases with different values of Y- For characteristics run at constant values of
UhlT the value of t will vary somewhat over the mass flow range. The value at
Performance with altered gas properties 193

the maximum efficiency flow will be suitable to define the whole characteristic
with acceptable accuracy.

Gases other than air

When gases other than air are considered the change of molecular weight, and
therefore of gas constant as well as specific heat have to be taken into
consideration. Table 24.1 illustrates the property variations for a number of
common gases.
Table 24.1
Properties of common gases

Gas MW G y a a/a*
Air 28.97 287 1005 1.4 340 1.0
Argon 39.94 208 518 1.67 316 0.93
CO2 44.01 189 652 1.29 275 0.81
Helium 4.00 2079 5277 1.66 994 2.92
Hydrogen 2.02 4116 14155 1.41 1293 3.8
Methane 16.04 518 2368 1.28 436 1.28
Freon 11 137.4 60.5 666 1.1 138 0.41

M.W. is the molecular weight


G is the gas constant in J/kg K
Cp is the specific heat at constant pressure in J/kg K at a temperature of 288 K
y is the ratio of specific heats = Cp/Cv at 288 K
a is the speed of sound at 288 K static temperature.

Note that in all cases G*(MW) = 8314 J/kg mole K which is the universal gas
constant

Saturated gases

When operating near to saturation, as in steam and refrigeration plant, gases tend
to depart significantly from the perfect gas law. In these circumstances the
constants given in the table, other than the molecular weight vary considerably
and it is necessary to use a Mollier diagram to determine the changes of state. A
series of these diagrams are given by Gresh (1991) for the gases most commonly
used in industrial processes.

The speed of sound

According to the kinetic theory of gases the value of y is 1.67 for a monatomic
gas, e.g. helium, and 1.4 for a diatomic gas, e.g. oxygen and nitrogen. For
194 Axialflow fans and compressors

polyatomic gases such as carbon dioxide, methane, and Freon 11 the value of y is
lower, as shown in the table. The speed of sound, a, is given by:

a « VCjfGt)

Since y varies only from 1.1 to 1.67 it has less influence on the sonic velocity
than the gas constant or molecular weight, which varies by a ratio of 68:1 for the
gases listed. Thus the speed of sound in hydrogen is more than 9 times what it is
in Freon 11. It can be shown that the pressure ratio achieved by a stage of given
blading is a function of the efficiency, the Mach number and the value of y and
Cp or the gas constant. Thus very high blade speeds and gas velocities are
required to produce the same pressure ratio with hydrogen compared to air. On
the other hand quite low blade speeds and gas velocities can produce the same
pressure ratio and Mach number with Freon 11. Advantage has been taken of this
to allow high Mach number research units to be run at modest speeds and
stresses by using Freon instead of air.

Compression of helium

The effect of using helium instead of air in a conventional stage illustrates the
problems which arise. Assume a stage designed for air to run at a blade speed of
300 m/s with a flow coefficient Va/U = 0.6 and a work coefficient AH/U2 = 0.4
and inlet whirl angle of 20°, at 288 K inlet total temperature. The temperature
rise across the stage would be 35.8°C and the total pressure ratio 1.45 for a
polytropic efficiency of 90%. The relative Mach number at inlet to the rotor is
0.9. If the stage is run at the same blade speed using helium the temperature rise
is only 6.82°C and the pressure ratio 1.055. The Mach number relative to the
rotor at inlet is only 0.3. While the stage efficiency might be somewhat higher at
this greatly reduced Mach number, further stages would be grossly mismatched
unless the annulus areas were suitably modified.
If, on the other hand, the stage were to be run with helium at the same Mach
number as for air the blade speed would require to be 860 m/s. This would be
impractical on account of the extreme mechanical stresses, but would produce a
pressure ratio of 1.5 compared with the ratio of 1.45 for air at the same Mach
number. Looked at the other way round, if a compressor were designed to use
helium it would be possible to obtain useful test data by running with air at a
suitably reduced blade speed.
The problem encountered when designing for a gas such as helium is that for
mechanically viable blade speeds the pressure ratio obtainable per stage is low
relative to that obtainable with air and other gases of greater molecular weight.
This has lead to experiments on designs featuring high values of AH/U2 and
tandem rotor blades such as in figure 24.2. Bammert and Staude (1979) have
shown successful results from a compressor having three stages of tandem rotors
Performance with altered gas properties 195

Tni]’"0
outk

Leading rotor,
outlet ,

Figure 2AJ2 Section of two row rotor and vector diagram


Adaptedfrom Bammert and Staude (1979)
with permission ofASME
with an inlet and outlet stage having conventional rotors. The basic vector
diagram was for 100% reaction and AH/U2 = 1.0 and Va/U = 0.75
approximately. The combined tandem rotors had a very low overall aspect ratio,
and it would be interesting to know whether single rotors of the same aspect ratio
would produce as good a performance.

Viscosity and Reynolds number

Another parameter which may affect the performance with different gases is the
Reynolds number. A table of kinematic viscosities for industrial gases is given
by Gresh and this shows air to have almost the highest value, while hydrogen is
one of the lowest with a value about half that for air. This means that the
Reynolds number will be two times greater in hydrogen for a blade of given
dimensions operating at the same velocity as in air.
25 Design for a domestic
ventilator

i
i

Specification

A domestic ventilator fan, suitable for a typical kitchen, requires a volume flow
of 0.2 m3/s. This would provide for about fifteen changes of air per hour in a
large domestic kitchen. The pressure difference from the room to the outside
would be negligible, and an acceptable velocity through the fan blade annulus is
10 m/s. The driving motor will be in the fan hub, and the discharge will be
directly from the annulus without any diffusion. The inner annulus diameter
must be large enough to accommodate the motor.

Overall parameters

The inlet total pressure, Pi, and the outlet static pressure, p2, will both be equal
to the atmospheric pressure, hence the total pressure rise will be:

AP = P j-P ,= p j + '/2pV2-P,

But P2 = Pi and so: AP = V: pVa2/cos2a

where P2 is the outlet total pressure, V is the outlet velocity, Va is the axial
velocity and a is the swirl angle of the exit flow. It is assumed that no outlet
guide vanes are required and the mean value of a is not more than 15°.

AP = 0.5* 1.225* 102/cos215° = 65.65 Pa

The work input per unit mass flow is given by:

AH = AP/pr] = 69 rrf/s2
Design for a domestic ventilator 197

A modest value of 80% is assumed for the blading efficiency to allow for losses
associated with motor support struts, relatively large tip clearance, low Reynolds
number and discharge losses. The volume flow, Q = AVa, and hence the power
requirement will be:

W = MAH = QpAH = QAP/n

= 0.1*65.65/0.8 = 8.2 watts

At this minimal level of power the motor and fan efficiency are obviously of
little concern and an induction motor would probably be of sufficiently excess
power to run close to its synchronous speed which may be chosen as 3000 rpm.

Blade speed and dimensions

Referring to figure 3.9 the constant radial value of AH/Va2 is 0.67, and in order
to avoid excessive outlet whirl angles which would make a row of outlet guide
vanes desirable, a hub value of Va/U of 0.6 is indicated as a maximum. The hub
blade speed is therefore:

Uhub = Va/(Va/U) = 10/0.6 = 16.7 m/s

For the selected rpm of 3000 we have:

Uhub = 3000nd/60 =157d = 16.7 m/s

Hence the hub diameter, d = 16.7/157 = 0.106 m

and the annulus areaA = Q/Va = 0.2/10= 0.02 m2 = ji(D2 - d2)/4

Hence the tip diameter, D = 0.192 m

and the diameter ratio,d/D = 0.106/0.192 = 0.552

These appear suitable dimensions which will accommodate the motor within
the hub diameter and also result in a suitable hub to tip diameter ratio. A value
much less than 0.5 for the diameter ratio could result in difficulties in blade twist
and radial matching of the design parameters. A diameter ratio greater than, say,
0.7 would lead to an excessive overall diameter and an unnecessary loss of
efficiency due to additional annulus surface area and a greater ratio of tip
clearance to blade height for the same radial clearance.
198 Axialflow fans and compressors

Blade selection

The rotor inlet and outlet angles for hub, mean and tip diameters are calculated
in table 25.1.

Table 25.1

Diam.mm 106 149 192


Um/s 16.7 23.5 30.2
Va/U 0.6 0.427 0.331
0tj0 59 66.9 71.7
AH/UVa 0.402 0.286 0.222
a 2° 51.7 64.1 70.34
tanOa 1.465 2.202 2.91
c° 52.7 64.0 70.1

<Xi is obtained from U/Va = tan Oi;


a 2 is obtained from AH/UVa = tana( - tan a.i
tana,,, = 0J(tanot| + tana2) and tan£ = tanOg, - 0.15

To obtain suitable blading using C5 circular arc cambers first consider the hub
section for incidences of -5°, 0° and +5° as in table 25.2.

Table 25.2

t° -5 0 +5
P»° 64 59 54
0° 22.6 12.6 2.6
P20 41.4 46.4 51.4
5° 10.3 5.3 0.3
1.1+.310 8.1 5.0 1.91
S/C 2.05 1.19 0.004

In the above table:


p, = Ct, - l
e = 2( p , - o
P: = Pi -8
5 = a 2 - P2
S/C = {5/(l.l+.319)}3

If the blade chord is to be approximately constant with radius, 0° incidence and


S/C =1.19 appear satisfactory values as this will give S/C = 2.16 at the tip, and
1.67 at 0.149m diameter.
Design for a domestic ventilator 199

From the relationships for the deviation:

8 = 02 - p2 = ot2 - ( C - 6/2)

and 8 = (1.1+0.316XS/C)'/I

the camber can be obtained for these values of S/C. This results in a camber of
11.9° at the dp and an incidence of -4.4°. While this would be acceptable, it
would be preferable to increase the chord and so bring the incidence closer to
zero. For a given number of blades this increase in chord would reduce the rado
of tip clearance to chord, and while the advantage in efficiency would be
unimportant the increase in pressure rise at stall could be an advantage.
For i= -3° the camber would be 9.2° and the S/C 1.84 at the tip and assuming a
linear variation of chord with radius the mid span SIC would be 1.52 which
would give a camber of 8.1°, with an incidence of -0.8°. The rotor relative air
angles 04 and 02 and blade angles pi and are shown in figure 25.1.

Diameter mm

Figure 25.1 Air and blade angles for ventilator

The blade sections derived above are of much lower S/C than would give
maximum efficiencies according to figure 4.3, however, this is unimportant in
this design because of the very low power consumption. A low number of blades
would be preferable to keep costs as low as possible, subject to the aspect ratio
being satisfactory from a stall margin point of view. Based on the mid span and
assuming an aspect ratio of unity the chord would be 43mm and for the S/C of
1.52 the blade spacing would be 65.4 mm. Hence the blade number would be
71*149/65.4 = 7.15. Seven blades would therefore be a suitable choice, and the
mid span chord would be 44mm.
200 Axialflow fans and compressors

Half speed operation

It is interesting to note that if the ventilator were operated at 1500 rpm it would
still give 7.5 changes of air per hour to the room, which may be adequate for
many purposes and would considerably reduce noise.
Alternatively if all linear dimensions were increased by V2 the annulus area
would be doubled and the volume flow would be the original 0.2 m3/s at the
same velocities as the basic size at 1500 rpm. In this case the rpm would require
to be 3000/V2 = 2121. Operated at 1500 rpm this size of fan would give a
volume flow of 0.2*1500/2121 = 0.14 m3/s. The larger size would of course lead
to increased material and manufacturing costs.
26 Design for an industrial fan

Specification

Volume flow = 5 m3
System duct diameter = 0.75 m
Duct system pressure drop = 500 Pa
Fan rotational speed preferably 1500 rpm

Preliminary considerations

A rotor plus outlet stator design is assumed, subject to this presenting no major
problems. Since the fan is to be installed in a system using ducting of 0.75m
diameter it would be convenient if the fan outer diameter was also 0.75m
diameter if this is found suitable. The fan will require to produce a total pressure
rise equal to the specified ducting pressure drop of 500 Pa plus the losses in
diffusing the annulus velocity back to the pipe velocity downstream of the fan.
The lower the fan annulus axial velocity, the lower these losses will be, and
therefore the larger the annulus area the better, which indicates a preference for
the lowest practical hub diameter for the fan. This may be limited by the
diameter of the driving motor, assuming it to be fitted within the fan hub. The
blading design may also place a limit on the minimum hub diameter, since blade
speed at a radius is fixed by the preferred rotational speed of 1500 rpm. It will be
assumed that the diffusion of the axial velocity downstream of the fan will have
a diffusion efficiency of 80%, i.e. the static pressure recovery is 80% of the ideal
corresponding to the area ratio. This is a conservative figure which should allow
for the wall boundary layer conditions entering the diffuser behind the fan. The
total to total pressure rise efficiency of the fan will be assumed as 85%, and this
will be taken as constant along the blade span. Again this is a conservative target
which allows for typical blade tip clearances, say about 2% of blade span.
202 Axialflow fans and compressors

Choice of diameter ratio

Assume a diameter ratio d/D = 0.7. The fan annulus area is given by:

A. = (it/4) D2! 1 • (d/D1) = (ji/4) 0J623 {1 • 0.49) = 0.225 m2

The fan axial velocity, Va = Q/A. where Q is the volume flow, and hence:

Va = 22.2 m/s

The velocity in the system ducting, Vd, is given by Q/A* where Ad is the cross
sectional area of the 0.75m diameter ducting.

V„ = 5»4/(it*0.752) =11.3 m/s

The ideal pressure recovery factor for the downstream diffuser is given by:

Cp, = 1 - (Va/VJ2 =0.741

Using the assumed diffusion efficiency of 80%:

Cp** = 0.8Cpi = 0.593

The loss coefficient is the difference Cpi - Cp**, and so the diffuser loss of total
pressure is:

AP = (0.741 - 0.593)*0.5* 1.225*V,2 = 45Pa

Hence the total pressure rise required of the fan is:

AP = 500 + 45 = 545 Pa

The fan work per unit mass flow is:

AH = APflpti) = 545/(1.225 * 0.85) = 523 m2/ sJ

The power required to drive the fan will be:

Power = pQAH = 1.225 * 5 * 523 = 3203 watts

A motor of 5 KW nominal rating would be suitable allowing for the motor


efficiency and a degree of overload towards stall.
Design for an industrialfan 203

From the above the value of AH/Va2 = 1.06 and from figure 3.9 this would be
acceptable provided the hub value of the flow coefficient Va/U is less than 0.7,
since higher values of VaAJ would lead to the rotor blade Cpi exceeding 0.5.
The blade speed at the hub will be:

U, = (1500/60)*** 0.75 * 0.7 = 41.2 m/s

Hence Va/U = 0.539 which would be acceptable.


Lower diameter ratios can be examined in the same way with the results
tabulated below:

Table 2 6 .1

D i/D 0 0 .7 0 .6 0 .5
Va 2 2 .2 1 7 .7 1 5 .1
AH 523 502 492
A H /V a 2 1 .0 6 1 .6 2 .1 6
V a /U 0 .5 3 9 0 .5 0 1 0 .5 1 2

Plotting of these values of AH/Va2 and Va/U on figure 3.9 shows that the value
of Cpi is just greater than 0.5 at a diameter ratio of 0.6, but greatly in excess of
this at a diameter ratio of 0.5. Thus a diameter ratio of 0.6 appears the most
suitable. It should also be noted from figure 3.9 that the value of a 3 is less than
45° at the hub, and as this will be the maximum outlet stator deflection, it will be
acceptable. Since AH/Va2 is constant along the blade span, it is also clear from
figure 3.9 that the Cp, values for the rotor will fall from hub to casing.

Rotor blade geometry

Table 26.2

Dia. m 0.45 0.55 0.65 0.75


Blade speed U m/s 35.3 43.2 51.1 58.9
Va/U 0.501 0.410 0.346 0.301
cti 63.4 67.7 70.9 73.3
AH/UVa 0.803 0.657 0.555 0.482
50.1 60.7 66.8 70.7
a3 38.7 33.3 29.1 25.0
Cpi (rotor) 0.513 0.399 0.310 0.244
S/C opt. (rotor) 0.488 1.514 2.31 2.91
Cpj (stator 04 = 0°) 0.391 0.301 0.237 0.179
S/C opt (stator). 1.58 2.39 2.974 3.5
204 Axialflow fans and compressors

In Table 26.2 the figures are obtained from the following relationships:
tan Oj = U/Va (for oto = 0°); tan a 2 = tan ct| - AH/UVa
tan a 3 = U/Va - tan a 2
Cpi (stator) = 1 - cos 2a3 (fora* = 0°);
Cpi (rotor) = 1 - (cos2ai/cos 2<X2)
S/C opt. = 9(0.567 - CpO

The rotor blade design is to give a work input of 502 m2/s2 at an axial velocity
of 17.7 m/s at all points along the blade span. While the optimum values of S/C
are noted in Table 26.2, it is found more practical to design blading within the
range of incidence from +5° to -5° for the hub and dp sections, and having
determined suitable values for S/C at these sections, to use a linear variation of
chord to fix S/C values at intermediate sections and then to determine the
corresponding values of camber.
For the hub section at 0.45m diam. oti = 63.4°; a 2 = 50.1°
tan £= tan a*, - 0.15 = 1.446; hence ?= 55.3°

Table 263

l -5 -2.5 0
Pi 68.4 65.9 63.4
e 26.2 21.2 16.2
P: 42.2 44.7 47.2
5 7.9 5.4 2.9
1.1+0.310 9.22 7.7 6.12
S/C 0.629 0.345 0.106

Table 26.3 clearly indicates an incidence between -5° and -2.5°. Before
finalising on a value the blade geometry at the tip is examined in the same way.
For the tip section at 0.75m diam:

a, = 73.3°; a 2 = 70.7°; hence C = 71.2

Table 26.4

i -5 -2.5 0
P. 78.3 75.8 73.3
0 14.2 9.2 4.2
P2 64.1 66.6 69.1
8 6.6 4.1 1.6
1.1 +0.310 5.5 4.0 2.4
S/C 1.73 1.08 0.296
Design for an industrialfan 205

The following relationships were used to evaluate the figures in the above
tables:

p, = a,-i; e = 2(P, -0 ; P2= P.-0; S = a2-p5; S/C = {8/(1.1+O.310))3

Preferably the chord at the dp should not exceed the hub chord by more than
20% for reasons of centrifugal loading, nor be less than, say, 2/3 of the hub value
in order to ensure a reasonable size of aerofoil and modest thickness/chord ratio.
If the chord were constant radially the dp S/C would be greater than the hub
value by the ratio of diameters:

D/d = 0.75/0.45 = 1.667

If the hub S/C = 0.629, then dp S/C = 1.048, for constant chord, and 1.572 if
tip chord were 2/3 of hub chord. From the table above it can be seen that both of
these values of S/C would give an acceptable incidence. The higher value of
1.572 is chosen since this is closer to the optimum S/C indicated in the previous
table. The blade height will be:

h = 0.5(D-d) = 0.15m

Hence an aspect ratio of 1.5 would result in a mid-height chord of 100 mm.
At the mid-height:

S/C = (1.572 + 0.629)/2 = 1.1

and S = 1.1 * 0.1, which determines the number of blades as:

n = Jt* 0.6/0.11 = 17.1


So 17 blades are chosen and the chords adjusted to give the selected values of
S/C, and the blade geometry can be calculated as in the table below.

Table 26J

Dia. m 0.45 0.55 0.65 0.75


S/C 0.629 0.94 1.258 1.572
OL\° 63.4 67.7 70.9 73.3
a 2° 50.1 60.7 70.9 70.7
c° 55.3 63.0 67.9 71.2
0° 26.4 17.1 13.7 13.0
p.° 68.5 71.6 74.8 77.7
i° -5.1 -3.9 -3.9 -4.4
p2° 42.1 54.5 61.1 64.7
206 Axialflow fans and compressors

The intermediate values of S/C are from a linear variation with radius. The value
of 6 is calculated from the relationship:

0 = {l.l(S /Q a+ C - «2}/{0J - 031 (S/C)2}

A C4 aerofoil on circular arc camber line has been assumed and a


thickness/chord ratio of 10% at the hub with a linear taper to 5% at the tip would
be suitable. Tip clearance of 1.5 mm, equal to 1% of blade height would be a
satisfactory standard; 3 mm would be the maximum acceptable.

The outlet stator

The hub and casing sections of the stator are considered in a similar fashion to
the rotor. For -5° of incidence at the hub a stagger of 14.1°, and a camber of
59.2° are obtained at an S/C of 0.506. Although a lower camber might be
preferable, this would result in an unattractively low S/C. At the outer diameter a
range of possible geometries are calculated as shown in table 26.6

Table 26.6
000

a 3 = 25°; 04 = 0°; hence £


11

1° -5 0 +5
p.° 30 25 20
e 50.4 40.4 30.4
p2° -20.4 -15.4 -10.4
5° 20.4 . 15.4 10.4
1.1+0.310 16.72 13.62 10.52
S/C 1.816 1.445 0.966

If the chord were constant radially the S/C at the outer diameter would be:

0.506*0.75/0.45=0.843.

This would require a design incidence in excess of 5° at the outer diameter. A


reduction of the chord towards the outer diameter is therefore indicated.
However, as the stator may be mounted from the casing, and be unshrouded at
the inner diameter, the chord would preferably be greatest at the outer diameter,
or radially constant. Assume a constant chord = 100mm, giving S = 96.5 mm at
the outer diameter. The number of stators will be:

n = it * 0.75/0.0965 = 24.4
Design for an industrialfan 207

There is a case for preferring prime numbers of blades on the part of some
designers, so let 23 blades be chosen, which will require a chord of 106 mm to
give S/C = 0.965 at the outer diameter. With constant chord the stator S/C at the
hub will be 0.579. The hub camber is calculated as 62.2° and the incidence as
*6.5° both of which are somewhat excessive. Thus unless an undesirable taper of
chord increasing towards the hub is adopted it is not possible to satisfy the
design within the limits preferred. An alternative possibility is to allow, say, 5° of
outlet whirl at the hub:

a 3= 38.7°; cu= 5° hence £ = 16.4°

C = 106mm; S = 61.5mm; and so S/C = 0.580

M(S/G)s+C -a««12.32

0.5 - 0.31(S/C)J = 0.2414

And so 0 = 51.0°; 0, = 41.9°; and \ = -3.2°

This solution reduces the camber significantly as well as giving a more


acceptable incidence. The small degree of outlet whirl at the hub should present
no problem and zero whirl can be retained over the outer half of the blade span.
The final blade geometry can be calculated as in the following table.

Table 26.7

Dia. m 0.45 0.55 0.65 0.75


0t4o +5 +2 0 0
a 3° 38.7 33.3 29.1 25.0
16.4 11.1 7.3 4.8
C mm 106 106 106 106
S mm 61.5 75.1 88.8 102
S/C 0.580 0.708 0.838 0.962
1.1(S/C)1+ C -Ok 12.23 10.08 8.34 5.89
0.5 - OJUS/C)1 0.243 0.224 0.207 0.192
0° 50.3 45.1 40.3 30.7
p.° 41.6 33.6 27.5 20.1
1° -3 +0.3 +1.6 +4.9

It is of interest to compare the above stator design with a design derived from
Howell's data. Allowing that Howell’s deviations are generally accepted as too
small, the deflections required can be increased by 2° to allow for this, i.e. the
design outlet angle will be assumed to be -2°. The nominal S/C at the hub is
208 Axialflow fans and compressors

indicated as 0.6, while at the casing it would be 1.6. The same considerations
concerning the radial distribution of chord would apply and so the S/C at the
casing would be based on a constant chord, giving S/C = 0.6 at the hub and 1.0
at the casing. The incidences would be chosen in the range +5° to -5°, so the
resulting blade would be very similar to the one detailed above, except that the
problem of keeping the incidences within the preferred range and avoiding
excessive camber does not emerge. This may be considered a weakness of the
Howell method.

The outlet diffuser

It was assumed in the above design that the tailcone diffuser downstream of the
fan would have a diffusion efficiency of 80%, i.e. the static pressure recovery
would be 80% of the ideal for the area ratio. Examination of figure 6 of Adkins
et al (1983) indicates that this could be achieved with a simple conical tail piece
to the hub having an included angle of 35° to 40°. Although the cone will have an
axial length of approximately of 1.5 times the hub diameter, the full static
pressure recovery will only be achieved at approximately twice this distance
from the entry to the diffuser. The same reference indicates that somewhat better
recovery of static pressure may be possible with smaller tailcone angles if the
extra length of cone is practical.
27 Design for a transonic
wind tunnel fan

Specification

A fan is required for a closed circuit wind tunnel with a maximum working
section Mach number of 1.4. A cooler is necessary in the circuit between the fan
and the working section to remove the temperature rise of the fan and for this
reason the fan inlet temperature will be maintained at 320K to allow for a normal
cooling water temperature. At the maximum working section Mach number of
1.4 the pressure losses around the circuit require a fan pressure ratio of 1.25
including diffusion to a Mach number of 0.3 after the blading. The working
section will be assumed to have a cross sectional area of 1 m2 and the pressure
ratio from the working section to the fan inlet is taken as 1.2.

Preliminary considerations

The specified pressure ratio should be readily obtained from a single stage and
since the inlet and outlet flows will require to be axial a zero inlet whirl stage is
attractive in that no inlet guide vane is necessary and no excessive whirl requires
to be removed after the rotor. Both of these considerations will save on pressure
losses and on capital cost; therefore, subject to the rotor inlet relative Mach
number being acceptable, this form of stage design is to be preferred. Subject to
other considerations the diameter ratio should be moderately low for best
efficiency, probably in the range 0.5 to 0.7. The axial Mach number at the rotor
inlet should not exceed 0.55 in order to avoid a requirement for excessive
diffusion from the fan blading to the specified downstream Mach number of 0.3.
The level of relative inlet Mach number at the rotor tip is preferably less than 1.2
and the level of the tip de Haller number should not be greater than 0.83 (Cpi not
less than 0.3) in order to avoid too lightly loaded a tip section which could cause
a low efficiency locally at the high relative inlet Mach number.
210 Axialflow fans and compressors

Selection of vector geometry

For the pressure ratio of 1.25 the required temperature rise is given by:
AT = (Ti/nXR<,‘lyr- 1)
The efficiency, T|, must allow for the downstream diffusion to 0.3 Mach number
as well as the blading efficiency. For this reason a first approximation will be a
value of 85% rather than the 88% to 90% which might be expected across the
blading alone.

Hence: AT = (320/0.85X1.25a2“ -1) = 24.8 °C


and the enthalpy rise: AH = Cp AT =1005 * 24.8 = 24933 m2/s2

For the zero inlet whirl stage AH/Va2 is constant radially and reference to
figure 3.9 indicates a value of unity may be a suitable first trial for this
parameter. If the maximum value for Cp-, is taken as 0.5 then the hub will have:

Va/U = 0.707 and AH/U2 = 0.5

From these values and the value for AH we obtain for the hub section:

U = 223.3 m/s and Va = 157.8 m/s

The static temperature of the inlet flow will be:

t = T - (Va7l2Cr) = 320 - (157.82/2010) = 307.6K

Hence the speed of sound of the inlet flow:

a = V(yG0 = V(1.4*287*307.6) = 351.6 m/s

and the inlet axial Mach No. = 157.8/351.6 = 0.449

The hub relative inlet Mach number to the rotor, M ^ = 0.778

Assuming a diameter ratio of 0.5 the tip blade speed will be:

223.3/0.5 = 446.6 m./s

T he relative inlet velocity to the rotor w ill be:


Design for a transonic wind tunnelfan 211

V,J = Vi* + U2 = 157.81+ 446.61

and so Vj * 473.7 m/s

Hence the relative inlet Mach number = 473.7/351.6 = 1.347

This is a higher Mach number than desired, so consider diameter ratios of 0.6
and 0.7 with the same hub section.

Table 27.1

Diameter ratio d/D 0.5 0.6 0.7


Tip blade speed U** 446.6 372.2 319
Rel., inlet Mach No.Mnini 347 1.15 1.012
Flow coefficient Va/Utj. 0354 0.424 0.495
Work coefficient AH/LT* 0.125 0.18 0.245
Rotor de Haller No. dHu, 0.884 0.849 0.809

Diameter ratio d/D

Figure 27.1 Relative inlet Mach number at rotor tip


for a range of diameter ratios

A diameter ratio of 0.6 would be acceptable, but before deciding on this it is


worth considering a lower value of AH/Va2. At a value of AH/Va2= 0.864 figure
3.9 indicates that a Cpi of 0.5 is not exceeded at any value of the work and flow
coefficients, so this would be a possible choice, taking Va/U = 0.9 and AH/U2 =
0.7 for the hub. Performing the same calculations as above gives the following:

Hub blade speed, uboba 188.7 m/s Axial velocity, Va = 169.8 m/s
Static temperature t = 306K Local sonic velocity, a = 351 m/s
Axial Mach No. = 0.484; Rotor inlet relative Mach No. = 0.723
212 Axialflow fans and compressors

Table 272

Diameter ratio d/D 0.5 0.6 0.7


Tip blade speed UUp 377.4 314.5 269.
Rel. inlet Mach No.Mnm 1.179 1.018 0.908
Flow Coefficient Va/Un- 0.45 0.54 0.63
Work Coefficient AH/U^p 0.175 0.252 0.343
Rotor de Haller No. dHup 0.856 0.812 0.77

Diameter ratio d/D

Figure 27.2 de Haller numbers for a range of AH/Va2values

A further range of possible designs for a value of AH/Va2 = 0.75 have been
added to figures 27.1 and 27.2. The hub design for these has been set at AH/U2=
1.0 at Va/U = 1.155. As the diagrams show this would allow the adoption of a
diameter ratio of 0.5 while remaining well below the maximum values set for tip
Mach number and de Haller number. The lower diameter ratio is attractive as
this generally tends to a higher efficiency. For this reason a design for 0.5
diameter ratio at AH/Va2= 0.75 is chosen for further examination.

The rotor blade

Hub section

U = 157.9 m/s and Va = 182.4 m/s, hence oti = tan ‘‘(U/Va) = 40.9°
Since AH/U2 = 1.0 with no inlet whirl velocity the rotor outlet angle, a : = 0°
For the vector mean angle, tan a„, = 0.5 (tan ot| + tan Ob) = 0.433

For the stagger angle, tan £ = tan - 0.15, hence £ = 15.8°


The highest value of S/C will be obtained with the most negative acceptable
incidence, hence assume i = -5°; which gives the blade inlet angle:
Design for a transonic wind tunnelfan 213

Pi = <X| - 1 = 40.9° + 5° = 45.9°

The camber angle 0 = 2 (P, - Q = 2(45.9° + 15.8°) = 60.2°

The blade outlet angle p2 = P ,- 0 = 45.9° - 60.2° = -14.3°

and the deviation angle 8 = - p 2 = 0° - (-14.3°) = 14.3°

From the relationship 5 = (1.1 + O.310)(S/C)I/3 the value of S/C is obtained as


0.377 This is an undesirably low value and, together with the very high value of
the camber, makes this design unattractive to the extent that a higher diameter
ratio, and a higher value of AH/Va2, would be preferable. From the data of
figures 27.1 and 27.2 a diameter ratio of 0.6 and AH/Va2 = 0.864 is chosen, with
hub values for AH/U2 and Va/U of 0.7 and 0.9 respectively. At the hub the
vector angles for the rotor will be:

a, = tan'1 U/Va = 48°

02 = tan*1 (U/VaXl-AH/U2)

= tan‘1(l/0.9Xl - 0.7) = 18.4°

tan a* = 0.5(tan 48°+ tan 18.4°) = 0.722

tan t = 0.722 - 0.15 = 0.572; £ = 29.8°

Alternatively tan £ = tan On- 0.213 and £ = 27°

Table 273

t -5 0 + 5
Pi = ai - 1 53 48 43
0 = 2(P ,-Q 46.5(52) 36.5(42) 26.5(32)
P2= P i - 0 63 ( 1) 11.5(6) 16.5(11)
8 = <*2 - p 2 11.9(17.4) 6.9 (12.4) 1.9 (7.4)
1.1+0.310 15.5(17.53) 12.4(14.1) 9.32(11)
S/C= [8/(l.l+O.310)]3 0.453(0.98) 0.172(0.68) 0.009(0.3) S/C

In the above table the figures in parenthesis are for £ = 27°. If the higher
stagger is to be used the choice must be for S/C = 0.453 since a more negative
incidence is undesirable.
The rotor inlet relative Mach number was calculated as 0.723. and so from the
Wright and Miller relationship:
214 Axialflow fans and compressors

0/ScosaIML= 0.155M, + 0.935 = 1.047

The design ai = 48°, therefore:

O/S =1.047*0.669 = 0.701

Using data for throat areas of DCA blades from the same source and taking t/C
= 8 % gives O/S = 0.60 and hence <Ximl= 55°. This is too far from the design inlet
angle of 48°, so try the lower stagger angle at i = 0°, 9 = 42° and S/C = 0.68. The
same calculation as before results in (Xjml = 47.5° which is satisfactorily close to
the design angle of 48°.

The rotor tip

At the outer diameter Va/U = 0.9*0.6 = 0.54 and AH/U2 = 0.7*0.62 = 0.252

a, = tan*l(U/Va) = 61.6°

02 = tan*1 (U/Va)(l - AH/U2) = tan 1 (1/0.54X1 - 0.252) = 54.2°

tan a,,, = 0.5(tanot| + tana2) = 0.5(tan 61.6° + tan 54.2°) =1.618

tan£ = tana,,, - 0.15 = 1.468; £ = 55.7°

The axial velocity is 169.8 m/s as calculated previously and the blade speed is:

U,ip = Va/(Va/U) = 169.8/0.54 = 314.4

t = T - Va2/2Cp = 320 - 1682/2010 = 306 m/s

and the sonic velocity, a = 350.6 m/s

V, = (Va2 + U2),/2 = 357.3 m/s

Hence the rotor inlet relative Mach No. = 357.3/350.6 = 1.019

A low speed design will be considered first but with the restriction that the
incidence should be in the range -2° < i < +2° at this level of Mach number. It is
then necessary to check the throat width of the blades, which for this marginally
supersonic design may be done using Wright and Miller’s relationship:

O/Scos ct|Mi = 0.155Mt + 0.935 = 1.093

<X| =61.6 and so the required O/S = 1.093*0.476 = 0.52


Designfo r a transonic wind tunnelfan 215

Table 27.4

I -2 0 +2 -0.5
Pi 63.6 61.6 59.6 62.1
c 55.7 55.7 55.7 55.7
6 15.8 11.8 7.8 12.8
fc 47.8 49.8 51.8 49.3
6 6.4 4.4 2.4 4.9
1.1 + 0.310 6.0 4.8 3.5 5.07
S/C 1.215 0.79 0.316 0.904
O/S 0.546 0.51 N/A 0.522
<*11*. 60 62.2 N/A 61.4

In table 27.4 the values of O/S axe calculated using the data for the throat areas
of DCA blades given in Wright and Miller (1991). A value of 4% was assumed
for the tip thickness to chord ratio. By interpolation (Ximl will approach the
design Oi of 61.6° when the incidence is -0.5°, and the right hand column has
been added to the table above to give the final design geometry chosen for the
rotor tip section.
The S/C at the tip for a constant rotor chord from hub to tip would be 1.133
and so the S/C of 0.904 would require the tip chord to be 125% of the hub chord.
The tip axial chord would be 79% of the hub axial chord and this would be
satisfactory.
Intermediate sections between hub and tip would be designed by the same
methods, using, as a first approximation, values of S/C to give a linear variation
of chord from hub to tip. This could be varied if necessary in a second
approximation but maintaining a smooth distribution of the chord.

Number of rotor blades

The number of rotor blades required will depend on the blade aspect ratio for
which a suitable value may be determined by specifying a value for the
equivalent cone angle of the mid span section of the blade. As discussed in
Chapter 6 this angle is given by:

tan a«, = [(l/nXh/CXS/C)],/2 ( cos,/2 a 2 - cosI/2 aO

Assuming a linear variation of the chord from hub to tip means that the mid-span
S/C is 0.792 at 0.8 of the tip diameter.

(Va/lDaid = (Va/UXub(d/D)/0.8 = 0.9*0.6/0.8 = 0.675

(AH/UJ)md = (AH/UJW d/D )J(0.8)2 = 0.7*(0.6/0.8)J = 0.394


216 Axialflow fans and compressors

Hence aj = tan*1(U/Va) = 56°

and a 2= tan 'OJ/VaXl - AH/U2) = 41.9°

Assuming the aspect ratio is 1.5 results in a«q = 4.04° which is as high a value
as desirable. The resulting number of blades is given by:

n = ic(h/C)(C/S)( 1 + d/D)/(l - d/D) = n*(1.5/0.792)(1 + 0.6)/(l - 0.6) = 23.8

Since 23 is a prime number this may be preferred to 24 for the number of rotor
blades.

Rotor diffusion factor

The diffusion factor is calculated as 0.377 using the incompressible form as


given in Chapter 4. Plotting this on figure 6.6 indicates the design at 80% of the
stalling diffusion factor with an end wall loss coefficient of 0.06 and e/C = 0.05
from which a range of acceptable aspect ratios for values of clearance to blade
height can be obtained as shown in table 27.5.

Table 27.5

e/h 0.01 0.02 0.03


h/C 5.0 2.5 1.67

Practical values for e/h would lie between 0.01 and 0.02, indicating aspect
ratios between 2.5 and 5.0 could be possible. The previous choice of aspect ratio
of 1.5 is therefore indicated as conservative but has the advantage of fewer
blades and therefore lower cost, as well as a more robust blade design.
Alternatively the value of e/C could be reduced to 0.02, giving a reduction of the
end wall loss factor to 0.43 at the same diffusion factor, which would now be
significantly less than 80% of the stalling value. For an aspect ratio of 1.5 the
clearance to blade height would be 0.0133 which is an achievable value.

The stator

The stator design will be of significantly lower stagger than the rotor due to the
zero outlet angle required. This makes it desirable to derive the stagger angle
from the relationship: tan £ = tan a.\ - 0.15 particularly to provide adequate stall
margin at the hub where the axial velocity will drop most rapidly as the mass
flow is reduced at constant speed.
Designfor a transonic wind tunnelfan 217

Stator hub

The inlet angle 03 = tan*1 [(U/Va) - tan a 2] = tan*![(lA).9) - tan 18.4] = 37.8°

The inlet velocity V3= Va/cos 09 = 168.9/cos37.8 = 213.8 m/s

The sonic velocity a = (1.4*287*344.8)1/2 = 372.2 m/s

The stator hub inlet Mach No. = 213.8/372.2 = 0.574

This is low enough that the design will be satisfactory if calculated on a low
speed basis. The stator inlet Mach number will fall with increasing radius for the
free vortex style of design adopted and so the same applies to all sections of the
stator. The base aerofoil section adopted is not critical and there will be no
performance difference between C4 and DCA blading. The former may be
preferred for its thicker leading edge on grounds of robustness.
The oudet angle 04 = 0®, and so £ = 13.4°. For 1 = -5° this results in 0 = 58.9°,
and S/C = 0.565. This is a high camber angle but is provisionally accepted.

Stator at outer diameter

The inlet angle is calculated as 24.95° and the stagger angle as 4.7°. It is
convenient to make the stator axial chord constant along the span and the S/C at
the casing is calculated on this basis.

At the hub S/Cn = (S/C)/cos£ = 0.565/cos 13.4 = 0.581

/.at the outer diameter S/C„ = 0.581/0.6 = 0.968

and S/C = 0.968 cos4.7 = 0.965

The deviation 8 * ou - = ou- R - (6/2)] = (1.1 + 0.316XS/C)W

from which an expression for the camber angle 0 is derived as:

e = [1.1(S/C)1/J +C - OU] / [0.5 - 0.31(S/C)1/3]

This gives 0 = 29.96° and therefore p 3= 19.68° and i = +5.27°

A less positive incidence may be considered desirable but this would lead to a
greater S/C and therefore a greater axial chord at the hub than at the casing. A
more negative incidence and higher S/C at the hub would avoid this, but only at
the expense of an unacceptably high camber angle. The design is therefore
retained as calculated.
218 Axialflow fans and compressors

Number of stators

The mid-span inlet angle is calculated as 30.26°, and the stagger angle is 8.07°.
Maintaining the same axial chord as for hub and casing sections gives S/C =
0.782. Proceeding as before results in a camber angle of 42.4° and an incidence
of +1°. The equivalent cone angle of the mid-span section is calculated as for the
rotor with the results tabulated below:

Table 27.4

h/c 1.5 2 2.5 3 4


a*, 2.3° 2.66° 2.98° 3.26° 3.76°
n 24 32 40 48 64

Even at an aspect ratio of 4 the equivalent cone angle is acceptable, however


the number of blades is large and a lower aspect ratio of, say, 2 would be
preferred. This would also permit lower thickness chord ratios than a higher
aspect ratio without endangering the mechanical strength of the blades. Values of
8% at the casing tapering to 6% at the hub would be suitable. The stator blade
number could be 31 if a prime number is preferred.

Stator diffusion factor

The diffusion factor of the mid-span section is calculated as 0.333. If the tip
clearance to blade height is set at 0.015 then, for the chosen aspect ratio of 2, c/C
= 0.03 and from figure 6.6 the wall loss parameter is 0.046.

Annulus dimensions

The working section of the wind tunnel is 1 m2 in cross section and at a Mach
number of 1.4 this corresponds to MVT/AP = 36.25. The pressure ratio from the
working section to the fan inlet is 1.2 and so at the fan inlet M^T/P = 1.2*36.25
= 43.5. The fan axial velocity has been determined as 168.9 m/s and the inlet
temperature is 320 K giving Va/VT = 9.44 and so from compressible flow tables
MVT/AP = 29.4. The flow annulus area at the fan inlet can now be obtained
from:

A,f = 43.5/29.4 = 1.48 m2

This figure would require to be increased to allow for the blockage of the inlet
boundary layers. The blockage factor is likely to be about 0.98 at the rotor entry
and so the geometric annulus area required at the rotor inlet will be:
Designfo ra transonic wind tunnelfan 219

At0 » 1.48/0.98 = 1.51m1

The inlet diameter ratio has been determined as 0.6 and so:

Outer diameter D = 1.733 m


Hub diameter d s 1.04 m.

A constant axial velocity has been assumed through the blading and in order to
maintain this some reduction of annulus area is necessary to counteract the
increase of density across the stage. A rising hub diameter is generally
considered preferable to a constant hub diameter as it assists the more heavily
loaded hub section by the centrifugal action of the increasing diameter through
the rotor. Although a tapered casing is usually less costly this should not be
significant for a single stage fan.
At the stator exit the total temperature will be 320 +24.8 = 348.8 K and
assuming an efficiency across the blading of 88% the pressure will be 1.26 times
the inlet pressure. At the stator exit:

Va/VTj = 168.9/V344.8 = 9.096

Hence, from flow tables:

MVTj/AjPi = 28.5 = (MVT,/A,PlXTi/r,)1/J(Pi/PiXA,/Ai)

and so the flow aiea required will be:

Ajp = (29.4/28.5X344.8/320)in *1.48/1.26 = 1.258 m2

The blockage factor at the stator exit will have fallen from 0.98 at the rotor
entry to 0.95 and so the geometric area required will be:

Aw = 1.258/0.95 = 1.324 m2
The inner diameter at stator outlet is therefore 1.148 m and the ratio d/D = 0.662.
At the rotor exit/stator inlet plane the flow angle at the mid-span has been
calculated as 30.26° and so we can write:

V/VT = Va/(cos a VTj) = 168.9/(cos30.26 V344.8) = 10.53

and the flow function M>/T/AP s 31.83 = MVT^AjpCos ajPj)

= (MVT./A.fP.XA.f/Ajfcos OjXTj/T .A ^ /P j)

31.83 = 29.4(1.48/Ajpcos 30.26X344.8/320)W(1/1.27)


220 Axialflow fans and compressors

A2f = 1.293 mJ

Ajo = 1.293/0.97 = 1.333 m2

Figure 27J Fan annulus diagram

Note that the pressure ratio across the rotor has been taken as 1.27. This
corresponds to an efficiency of 92% across the rotor compared to 88% across the
complete stage. A blockage factor of 0.97 assumes a smaller increase of
blockage across the rotor than the stator.
The inner diameter at rotor trailing edge will be 1.148 m and the annulus
diagram would be as indicated in figure 27.3 above.

Design speed

The blade speed at the rotor hub at inlet has been calculated as 188.7 m/s and
this is at a diameter of 1.04 m. The design rotational speed is therefore:

N = 60*U/(rcd) = 60*188.7/(k*1.04) = 3465 rpm

Should the inlet temperature to the fan vary this rotational speed will require to
vary as the square root of the temperature to maintain the design pressure ratio
and mass flow function.

Fan efficiency

For design purposes efficiency values based on experienced judgement have


been used in the calculation of the fan blading. It is appropriate to check these
values by direct estimation of the losses. Three methods will be illustrated to
indicate the variations that arise in the assessment of pressure losses based on the
mid-span blade geometry, aspect ratio and end clearances.
Design for a transonic wind tunnelfan

HowelVs method

For the rotor.

Cop = 0.018; C * = 0.02*0.792/1.5 = 0.011

Q. a 2*0.792(tan56 - tan41.9)cos50 = 0.596

Cu, = 0.018*0.5962 = 0.0064

Hence CD= 0.018 + 0.011 + 0.006 = 0.035

O, = 0.035cos256/(0.792cosJ50) = 0.052

AIWpU 2 = 0.5(D) (V,/U)2 = 0J*0.052/sinJ56 = 0.038

For the stator

Similar calculation as above gives ©3 = 0.043

AIWpU 2 = 0J*0.043(Va/Ucosa3)2 = 0.0132

For the stage:

AP/pU2 = AHA;2 - (APk>JpU2U » - (A PtoJpU 'W

= 0.394 - 0.038 - 0.013 = 0343

Hence efficiency = 0343/0.394 = 0.87

Leiblien method plus Wright and Miller end wall loss

For the rotor

Diffusion factor, Dp = 0.379

©p » 2(Dp + 0 .3)/100(S/C)cosa2 = 0.023

Aspect ratio = 1.5 and e/h = 0.015

Hence e/C = (e/hXh/Q = 0.0225

From figure 6.6 the end wall loss parameter is 0.045


©Ew(h/CXV1/V2)2 = 0.045 and so (Dew = 0.017
222 Axialflow fans and compressors

Total loss coefficient = 0.023 + 0.017 = 0.04

APioJpU* = 0.04/(2sin256) = 0.029

For the stator.

By similar calculation for the stator APioo/pU2 = 0.01

Hence = (0.394 - 0.029 - 0.01)/0.394 = 0.90

McKenzie method

Because this method is based on a series of low speed compressor tests the
efficiencies obtained represent the total losses without the calculation of separate
profile and secondary losses. They will however require correction for Reynolds
number as well as Mach number if necessary.
Rotor Cpi = 0.436, and S/C = 0.792

From figure 4.4 GJ = 0.044

APioJpU1 = 0.044/(2sinJ56) = 0.032

Stator Cpi = 0.254, and S/C = 0.782

From figure 4.4 CJ3 = 0.085

APtoJpU2 = 0.085(Va/U)J/2cosJctj = 0.026

= (0.394 - 0.032 - 0.026)/0.394 = 0.852

The tests on which this figure is based were made at a Reynolds number of
1*10*. Correcting this to a Reynolds number of 3*105by the method described in
Chapter 10 gives an efficiency of 0.872 which is very close to that given by
Howell’s method, but almost three points greater than the Lieblein plus Wright
and Miller method.
The actual Reynolds number of the fan operating at an inlet total pressure of
101.3kPa and 320K inlet temperature is 4.6* 106 and when the efficiency is
corrected to this condition an efficiency of 0.909 is obtained. The upper critical
Reynolds number as given by Miller( 1977) (see Chapter 10) is calculated as
5.7* 106 for a surface finish typical of forged blading and since this is greater
than the operating value at standard atmosphere inlet pressure there would be no
gain of efficiency from polishing the blades. Were the operating pressure to be
raised by more than 25% polishing would become progressively more desirable.
Design for a transonic wind tunnelfan 223

Outlet diffiucr

If the constant diameter casing is continued downstream of the blading and a


conical exhaust fairing is fitted to the rear of the hub the flow will be diffused to
a Mach number which is likely to meet the specified 0.3 or less.

Axial velocity into the diffuser = 169.8m/s atT = 344.8K so V/VT = 9.14

From tables MVT/AP = 28.65 and MN= 0.466

The diffuser area ratio from the outlet annulus to the tip of the cone will be:

A3/A2 = DVOD2 - d2) ■ 1.7332/(1.7332 - 1.1482) = 1.78

Assuming a 2% loss of total pressure, the diffuser exit flow function will be:

MVt/AjP3 = 28.6/(0.98*1.78) = 16.4

From flow tables = 0.244

This is lower than the specified 0.3 Mach number and is acceptable if the loss
in overall efficiency is acceptable. Adkins et al (1983) present data for this type
of diffuser and indicate at their figure 6(d) that for a similar area ratio a static
pressure recovery factor of 0.58 can be achieved with an included cone angle of
25° at a length of three times the diffuser inlet hub diameter downstream of the
diffuser entry.
For incompressible flow the loss coefficient can be written as:

m = Cpi- Cp» = [1 - (A j/A j)2] - 0.58 = 0.104

From tables P2/P2 = 1.16; so taking 0 = APkm/(Pj - pt) we obtain:

AP10J P , = 0.104(1 -1/1.16) = 0.014

and P2/P 1 = 1 - (APioJP,) = 0.986

This is a marginally smaller loss than assumed above, so no iteration is


required. The work by Adkins et al used a uniform flow at diffuser entry, which
the fan will not provide. To allow for this assume a diffuser pressure ratio of
0.98. Using the design temperature rise of 24.8° and the calculated efficiency of
0.909 across the blading, die pressure ratio to the end of the diffuser becomes
1.26, which is marginally greater than specified.
224 Axialflow fans and compressors

Power requirement

At the fan inlet MVT/P = 43.5 and therefore for P = 101.3kPa and T = 320K the
mass flow will be 246 kg/s.

Power W = MAH = 246* 1005*24.8/1000 = 6139.6 kW

The non dimensional group for power is W/PVT which would have a value of
3.388. The power requirement at other values of the fan inlet total pressure and
temperature can be readily obtained from this number.

Operation at higher tunnel Mach numbers

Closed circuit wind tunnels designed for a maximum working section Mach
number in excess of 1.0 have a variable throat area upstream of the working
section followed by a divergent nozzle giving a supersonic acceleration from
Mach 1 at the throat to the supersonic operating Mach number. Because the
throat area is reduced as the working section Mach number is increased above
1.0 the flow function is a maximum at Mach 1, or slighdy greater. The tunnel
circuit pressure losses continue to increase with increasing Mach No. and so the
shape of the operating line on the fan characteristics presents a problem due to
the fan inlet flow function falling while the pressure ratio rises for tunnel Mach
numbers in excess of 1.2, as illustrated in figure 27.4.

Tunnel
Mach No.
1.5 r

& 1* .

0.3

20 30 40 50
Fan inlet flow MT^/P

Figure 27.4 Wind tunnel fan operating line

Possible solutions

A number of solutions to the matching of the fan operating line are possible.
Design for a transonic wind tunnelfan 225

1. A single stage fan is used for Mach numbers up to 1.2 and an additional
stage is added for operation at higher Mach numbers. This entails a time
consuming change over when moving between the lower and higher ranges
of Mach number.

2. For Mach numbers in excess of 1.2 a proportion of the fan discharge is


recirculated to the fan inlet so that the fan mass flow is in excess of the
working section flow. It transpires that about one third of the fan discharge
flow has to be recirculated at a Mach number of 1.8. This is equivalent to a
fan mass flow 50% greater than the working section flow and therefore the
power requirement would also be approximately 50% greater than the
minimum to meet the duty if the fan could be matched for the 1.8 Mach
number case. Although a single stage design would be possible, the fan
would most probably be a two stage design with a design pressure ratio of
IS to meet the 1.8 Mach number requirement
28 Design for an industrial
compressor

Introduction

In contrast to aero-engine compressors, industrial plant compressors are


generally of more conservative design. They often employ the same blade design
in several or all stages, merely cropping the blade length to suit the diminishing
annulus height towards the rear. Constant hub diameter is often preferred for this
reason, although some designs use a rising hub line through the first few stages.
Where the drive is by steam or gas turbine the design rpm may be selected to suit
the compressor design, but in some cases electric drive at synchronous speed is
preferred, i.e. 3000 or 3600 rpm depending on the frequency of the supply.
These are most often constant speed machines and for this reason they feature a
number of stages of variable stagger stators in order to allow variation of the
mass flow at a fixed pressure ratio. Most manufacturers offer a range of sizes to
suit various requirements, using as many common parts as possible. To provide a
wide range of mass flows it is necessary to have a range of gear drives since the
blade speeds will be similar at all sizes, requiring rotational speeds inversely
proportional to the diameter.

Design specification

A series of compressors is required to provide a range of mass flow and pressure


ratio. The same rotor blades with tip cropping are to be used in as many stages as
practical with constant hub diameter. The largest size will be that obtainable with
3000 rpm. Smaller mass flows will be obtained by removing front stages, and
beyond what is then obtained the basic design will be scaled down and driven
where necessary by a step-up gear drive.
Design for an industrial compressor 227

General considerations

Because the hub section blading of all stages will be the same it is important that
the blading design at this diameter is moderately loaded, both in terms of
parameters such as de Haller number, diffusion factor, or lift coefficient as well
as AH/U2. The latter in particular will require a modest diameter ratio at inlet and
a value of 0.6 would be a suitable first estimate. The first stage should have a
moderately high tip speed, and a high subsonic relative Mach number to the rotor
tip. Since the first few stages will require variable stagger stators an inlet guide
vane will be necessary. This suggests that some positive inlet whirl should be
incorporated to reduce the relative Mach number onto the rotor, at least towards
the outer diameter. A modem industrial design organisation would employ both
an axisymmetric throughflow program and a blade to blade computer program to
conduct the design. For illustration purposes here, the design is to be conducted
without the benefit of these methods. The blading will therefore be specified
either as DCA sections or C4 aerofoils with circular arc camber lines. Without a
throughflow program it is necessary to base the vortex flow on some simple but
reasonable approximation. Simple equilibrium of the vector mean velocities is
chosen for this reason and a ‘constant reaction’ obeying the equation:

Vwm= (l-R)U

From this and the simple equilibrium equation a relationship for the radial
distribution of axial velocity was derived in Chapter 8. The lowest suitable value
of R, as indicated on figure 8.7, would be 0.6, in order to avoid too steep a fall of
axial velocity towards the outer diameter. Values over 0.75 are likely to require
too little stator outlet whirl to give a useful reduction of the rotor tip inlet Mach
number.

Design parameter selection

Indications have been given above for some of the design parameters. To initiate
the design process it is necessary to make further choices for specific values of
some of the salient parameters. These are best based on an experienced
judgement, nevertheless, some may be found inappropriate, and changes may
have to be made as the design proceeds. Rotor inlet relative Mach number at the
tip of the first stage is chosen as 0.95, and the tip blade speed as 375 m/s. The
inlet hub/tip diameter ratio is selected as 0.6, and this gives a mean blade speed
of 0.8*375 = 300m/s.
At the mid radius a flow coefficient of 0.5 is selected. This should avoid
excessively high values at the hub and excessively low values at the tip. The
mean radius axial velocity will be 150m/s and, taking this as an approximation to
the mass average value, means an axial Mach number at inlet of approximately
0.45. This is a modest value and could be raised to 0.5 without difficulty, while a
228 Axialflow fans and compressors

limit of 0.55 would be considered acceptable for an industrial machine. Based on


the mid-span VaAJ of 0.5, a value for AH/U2 of 0.3 can be chosen, based on
figure 3.7. A somewhat higher value, say 0.35, would be possible for the mid­
span but this would not allow for any increase of Cp; towards the hub. The value
of AHAJ2 at the outer diameter will be:

AHAJ2 = 0.3 *0.82 = 0.192

Massflow

The mass flow of the compressor is properly found by integration of the axial
velocity along the blade span at the first rotor inlet plane, for present purposes
the axial velocity at mid span will be assumed to give a close approximation to
the mass average axial velocity. If a mass flow of 100 kg/s is specified at inlet
conditions of 101.3 kPa and 288K the inlet flow function will be:

m VTi/P,= 1CKW288/101.3 = 16.75

and Va/VT, = 150/^288 = 8.84

and hence MVT|/A|P| =27.9 (from tables)

Note that there is a small error in assuming the total temperature of 288K in
association with the axial velocity when there is a tangential velocity component
at the outlet of the IGV, but this negligible in terms of VT.
The required annulus area at rotor inlet, A| = 16.75/27.9 = 0.600m2. This
effective flow area requires to be increased due to the blockage of the annulus
boundary layers and blade wakes, which at the IGV trailing edge will typically
require a blockage factor of 0.98, resulting in a geometric or ‘metal* area of
0.6/0.98 = 0.613 m2. The diameter ratio has been chosen as 0.6 which results in
an outer diameter of 1.104 m, and a hub diameter of 0.662 m, giving a blade
span at the leading edge of 0.221 m and a mean diameter of 0.883 m.

Rotational speed

The blade speed at the outer diameter is 375m/s so the rotational speed can now
be calculated as:

N = (60U)/(tcD) = (60*375)/(Tt*1.104) = 6487rpm

Derivation o f the vector geometry

Based on the equations developed in Chapter 8 for equilibrium of the vector


mean velocities we have:
Design for an industrial compressor 229

V a ,2 = V a „ 3 - 2 ( 1 - R ) 2(U ,! - U „ J)

From this the most suitable value of R can be derived by examining the tip
section parameters.

For Um= 300: U, = 375 m/s and Van, = 150 m/s

Va»2 = 1502 - 2(1-R)2 * 50625

For R = 0.75:

Va, = 127.2 m/s, and AH/2U = 0.192 • 375/2 = 36 m/s

Vw0 = (l-R)U - AH/2U = 93.75 - 36 = 57.75 m/s

Oo = tan',(Vw0/Va) = tan*,(57.75/127.2) = 24.4°

Vo= Va/cosa= 139.7m/s

and so the static temperature is:

t = 288 - 139.72/(2 * 1005) = 278K

The sonic velocity, a = (tGi),/2 = (1.4*287*278),/2 = 334.2m/s

The relative rotor inlet velocity Vj = Va2 + Vwi2 hence:

Vw, = U - Vw0 = 375 - 57.75 = 3l7.3m/s

From which Vt = 341.8 and the corresponding Mach number is:

Mn = 341.8/334.4 =1.022

This is ahigher Mach number than desired, so a lower value of reaction is


required in order to increase the inlet whirl leaving the IGV, say R =0.7.
Carrying out the identical calculations again for this value of R gives the
following rotor tip values:

Va = 115.7m/s; cto = 33.4°; MN= 0.957

These are satisfactory, so conditions are calculated for other radii using R = 0.7
and are given in table 28.1. The axial velocities are taken as constant through the
stages to suit the use of common blading throughout The values of Cpi are
incompressible values. They are low enough that the slightly higher
230 Axialflow fans and compressors

compressible values should present no problem. In practice values of the above


parameters would be derived for two or more additional intermediate sections

Table 28.1
Parameters for R =*0.7

r/rijp 0.6 0.8 1.0


Va 172 150 115.7
Oo 2.5 16.7 33.5
mn 0.837 0.889 0.957
AH/U2 0.533 0.3 0.192
Va/U 0.764 0.5 0.309
a, 51.7 59.5 68.8
a2 29.6 47.7 62.9
0t3 36.5 42.0 52.2
Cpi(RTr) 0.492 0.431 0.37
Cpi(St’r) 0.353 0.398 0.459

Rotor blade geometry

The low speed data used for the industrial fan design of Chapter 26 can be used,
but with the addition of a check on the margin from choking at the operating
Mach number. As indicated by Wright and Miller (1991) the inlet angle for
minimum loss increases with Mach number. For this reason the stagger angle
will be derived from:
tan £ = tan On, - 0.213

rather than using the constant 0.15. The use of the higher constant results in a
higher incidence and lower stagger angle for the same S/C. This is more likely to
avoid the blade passage being choked. For the hub section:

tan a™= 0.5(tan 51.7° + tan 29.6°) = 0.917

From which the stagger angle, £ = 35.2°

For i = 0°; (3, = 51.7°; 0 = 33; p2 = 18.7° and S/C = 0.89

This S/C is too large for Cpj = 0.492, which has an optimum S/C = 0.675.

After some trials, try: i = +1.5°, which gives 0 = 30° and S/C = 0.738

Using Wright and Miller's data for the throat width of DCA profiles gives:
Design for an industrial compressor 231

0/C = 0.473 and O/S =0.641

0 /Scos<x,ml = 0.837 * 0.155 + 0.935 = 1.031

Hence cos (Ximl = 0.641/1.031 * 0.622, and (Ximl = 51.6°

This appears satisfactory, so proceed to check the maximum Mach number from
figure 9.4 using O/Scos <X|= 1.031. This gives a Mach No. of 0.9 which is
satisfactory for the operating Mach No of 0.85.

For the mid-span section: tan On = 0.5(tan 59.5° + tan 47.7°) = 1.398

tan C = 1.398 - 0.213 = 1.185; and C = 49.8°

O/Scos a, = 0.155 * 0.889 + 0.935 = 1.072

For / = 0°: 0=19.4° and S/C = 1.219

Assuming t/C = 0.06, Wright and Miller's data for the throat areas of DCA
aerofoils gives: O/C = 0.7186, and so O/S = 0.5895, and (Ximl = 56.7°, which is
too low compared to the design angle. After some trials it is found that an
incidence o f+1.5° gives a minimum loss angle of 59.4°. The blade geometry is:

Pi = 58°; 0 = 16.4°; S/C = 0.96

For the tip section:

tan On, = 0.5(tan 68.8 + tan 62.9) = 2.266

tan C= 2.266- 0.213 = 2.053; C = 64°

O/Scos <x )Ml = 0.957 * 0.155 + 0.935 = 1.083

For <Xide» = <Ximl O/S = 0.392

After trials of a number of incidences around zero this value of O/S is satisfied
at:
i = -1.5°, assuming t/C = 0.04.

The resulting blade geometry is:

P, = 70.3°; 0=12.6°; S/C =1.12


232 Axialflow fans and compressors

The rotor blade geometry is plotted as air and blade inlet and outlet angles in
figure 28.1. In this form the incidence, i, and the deviation, 5, can be readily
seen. Alternatively, the blade geometry can be presented as camber, 6, and
stagger, £ as in figure 28.2, which, together with chordal dimensions is more
suitable for the preparation of manufacturing data.

Radius ratio r/r*

Figure 28.1 Rotor inlet and outlet air and blade angles

Radius ratio r/r,

Figure 28.2 Rotor blade camber and stagger angles

Table 28.2

r/r, 0.6 0.8 1.0


S/C 0.738 0.96 1.12
C/C hub 1.0 1.02 1.098
C° 35.2 49.8 64.0
C * x /C « hob 1.0 0.810 0.60
Design for an industrial compressor 233

To avoid stress problems it is necessary that the chord of the rotor blade should
not increase excessively from hub to dp. It is also desirable that the axial
projection of the chord should not reduce excessively towards the tip, causing
the leading and trailing edges to lean too far from the perpendicular in the axial
elevation. A radial distribution lying between constant true chord and constant
axial chord is generally acceptable. The values calculated in table 28.2 are
satisfactory in these respects.

Aspect ratio and number of rotor blades

Using the equivalent cone angle as described in Chapter 6, assuming an aspect


ratio of unity, the mid-span blade geometry and design air angles give a cone
angle of 3.4°. An aspect ratio of one would be an unusually low value for a first
stage, and in the case of a compressor intended for constant speed operation this
is particularly so, since the problems of low speed stall are not a major
consideration. The intention to use variable stagger stators on the early stages, to
give increased range of mass flow, also alleviates any possible stall problems.
The aspect ratio will fall through the stages as the same blades are to be cropped
for succeeding stages. If an aspect ratio of 2 is chosen for the first rotor, giving
an equivalent cone angle of 4.8°, the aspect ratio will fall to about unity at the
5th or 6th rotor. Later stages may then adopt smaller chord rotors, staiting with
an aspect ratio of 1.5 or thereabouts.
The number of blades is given by:
n = 2rcr/S

where n is the number of blades, r = radius, and S = blade pitch.


For the mid-span radius:

S/rm= (S/C)(C/hXh/rro)

and h/rm= (r, - rh)/rm= (1 - 0.6)/0.8 = 0.5

Hence for mid-span, S/r = 0.96 * 0.5 * 0.5 = 0.24

and n = 2n/0.24 = 26.2; say 26 rotor blades.

Rotor blade chord

The number of rotors has been determined as 26 and the mid-span S/C as 0.96 at
a stagger angle of 49.8° and so the true and axial chords at mid span can be
found as:

S = n D/n x n*0.883/26 = 0.107m


234 Axialflow fans and compressors

C = S/(S/C)« 0.107/0.96 = 0.11 lm

C» = Ceos? s 0.11 lcos 49.8 = 71.6 mm

It has also been determined that C * / C * * * b = 0.81, hence Cu m 5 88.4 mm


and similarly C„ ap = 0.6Cn hob s 53 mm.

Stator blade geometry

Because a reaction of 70% has been chosen the stator blade will be of
considerably lower stagger than the rotor, and the Mach number at inlet to the
stator will be lower than at rotor inlet. For these reasons a C4 profile with a
circular arc camber line may be preferred to a DCA profile, and the stagger angle
will obtained from:

tan £ = tan a -0.15

The outer section has the highest value of Cpi = 0.459, corresponding to an
optimum S/C of 0.972. However it is found that this S/C results in an incidence
more negative than -5°, so a design incidence of >5° is chosen. The stagger angle
from the above equation is 39.5°, resulting in a camber angle of 35.4° and S/C =
0.91. A convenient arrangement for the stator design is a constant axial chord
radially, if this is acceptable on all other considerations. At the mid-span (r/rt =
0.8) the stagger angle is 24.2° and the S/C for constant axial chord is obtained as
follows:

At r/rt = 1.0; S/C„ = 0.91/cos39.5° = 1.179

Hence at r/r, = 0.8; S/C„ = 1.179 * 0.8 = 0.943

Thus we obtain: S/C = 0.943 * cos24.2° = 0.86

By manipulating the relationships for deviation an equation for camber at a


known S/C can be obtained:

8 = a 2 - p2 = a 2 - C + 9/2 = (1.1 + O.310XS/C)1

From which: 0 = [1.1 (S/C)1+ £ - a 2] / [0.5 - 0.31 (S/C)2]

This gives: 0 = 41.6°; p 3 = 45° and i = -3°


Design for an industrial compressor 235

Mid-span
at stage Mid-span
10 outlet at stage
1 inlet O/D at

Figure 283 Axial velocity variation with radius

Since the incidence is in the range -5° < i < +5° this is an acceptable result For
the hub section the above procedure can be repeated with the following results:

S/C = 0.687; 6 = 53.3°; p3 = 40.3° and i= -3.8°

Stator aspect ratio and numbers

The equivalent cone angle is derived as for the rotor and an angle of 4.5° is
obtained at an aspect ratio of 1.5. The ratio of blade span to mean radius is:

h/r. = (rc -rh)/rm= [l- = (1 - 0.6)/0.8 = 0.5

where rc is the casing radius, rh is the hub radius and rmis the mid-span radius

S/r = (S/CXC/hXh/r) = 0.86*(1/1.5)*0.5 = 0.287

n = 2ror/S = 2*A).287 = 21.9

A choice of 23 stator blades would be appropriate if a prime number is preferred.


As determined previously the mid-span of the stator is to have S/C« = 0.943
and so C« can be found:

S = itDJn = **0.883/23 = 0.121

CM= 0.121/0.943 = 0.128m


236 Axialflow fans and compressors

Annulus diagram

The same rotor blades are to be used for several stages with appropriate cropping
of the tip and the stators for several stages will also be identical except for
modification of their span to suit the falling casing diameter through the stages.
The blade chords having been determined, the annulus diagram can be drawn
when annulus areas at each stage have been calculated and blade axial spacings
decided upon.
Because of the tapered casing with a constant hub diameter the mid-span radius
will fall through the stages and since the axial velocity rises towards the hub the
mid-span axial velocity will rise from stage to stage. From figure 28.1, as a first
estimate, assume that the axial velocity at mid-sjjan of stage 10 exit is 160 m/s.
The overall enthalpy rise will be 10*27000 m'/s2 and assuming a mean Cp of
1.01 gives an overall temperature rise of 267.3°C. For a polytropic efficiency of
88% and taking a mean value for y as 1.395 we obtain an overall pressure ratio
of 7.67. A first approximation to the area required at the tenth stage exit can now
be estimated.

Va/VT = 160/(288 + 267.3) = 6.79 and so MVT/AP = 22.4

MVT/P = 100*(288 + 267.3)l/:/(7.67* 101.3) = 3.03

Hence the flow annulus area, AP = 3.03/22.4 = 0.13S m2 and the geometric or
‘metal’ annulus will be 0.135/0.88 = 0.153 m2 where 0.88 is the blockage factor
at this station. The hub diameter is known to be 0.662 m and so the outer casing
diameter is 0.795 m and the mid-span diameter is 0.729 m and the ratio r jr h =
0.729/0.662 = 1.101. From figure 28.3 a mean axial velocity of 166 m/s is
indicated and this value of Va is used for a second iteration to obtain a final
value of the annulus dimensions. A close approximation to the mean axial
velocity at each stage can now be made and the calculation can be repeated to
specify the annulus area at each stage. The only missing quantity is the value of
the blockage factor at each stage. Typical values for the blockage factors through
the stages are given in table 28.3

Table 28.3

Stage No. 1 2 3 4 5 6
Blockage 0.98 0.95 0.92 0.9 0.88 0.88

The blockage factor is taken as constant for stage 5 and beyond.


Design for an industrial compressor 237

Table 28.4

Rotor inlet 1 2 3 4 5
Pn*l/Pa 1.317 1.288 1.264 1.243 1.225
P»+l/Pl 1.317 1.696 2.144 2.665 3.265
Csg. Dia. m 1.104 1.044 0.997 0.956 0.923

Rotor inlet 6 7 8 9 10
PIH-l/P0 1.210 1.197 1.185 1.175 1.16
Pm-l/Pl 3.953 4.732 5.610 6.589 7.665
Csg. Dia. m 0.891 0.865 0.841 0.824 0.807

While the value of AH is constant for each stage the value of the stage
temperature rise diminishes with the rising value of specific heat, resulting in a
temperature rise of 25.9°C on the tenth stage compared to 26.9°C on the first
stage. The value of y also falls with increasing temperature, but for design
purposes the stage polytropic efficiencies are assumed constant It should be
appreciated that the blockage factors and the efficiencies are approximations
based on general experience. The beneficial effect on efficiency of falling Mach
number with increasing temperature may offset the efficiency reduction to be
expected from increasing diameter ratio towards the rear of the compressor. The
result of these assumptions gives the values for the stage pressure ratios and the
cumulative pressure ratios in table 28.4. Note the falling stage pressure ratio
towards the rear stages, while the cumulative pressure ratio increases more
rapidly. The casing diameters are for the rotor inlet stations. At stage 10 outlet
the casing diameter is 0.793m.

Variation of stage numbers

Clearly by having a smaller or greater number of stages the pressure ratio and the
design mass flow can be varied over quite wide limits. By removing stages from
the rear the pressure ratio can be reduced at constant mass flow and by adding
further stages of the same basic design the pressure ratio could be increased.
Removing stages from the front will reduce the mass flow but to maintain the
same stage matching the new front stage should be operated at the same Mach
number as in the original design if the stage by stage annulus dimensions are
retained. This is obtained by reducing the design rotational speed so that the
value of N/VT on the new front stage is the same as for that stage in the original
design. For example suppose two stages are removed from the front
For stage 3 in the original design Tin = 341.8 K; and H*. = 6487rpm

N/VTin = 6487/^348.1 = 350.9


238 Axialflow fans and compressors

Therefore the new design speed = 350.9*V288 = 5955 rpm. At this speed the
pressure ratio of each stage would as tabulated above and the design mass flow
can be obtained from the value of the non-dimensional flow function at stage
three inlet of the original design.

Stage three inlet M>/T/P = 100*^341.8/(101.3*1.696) =10.76

Hence for stage 3 as inlet stage M =10.76* 101.3/V288 = 64.2 kg/s

This figure requires to be adjusted for the fact that the blockage factor of stage
3 will increase from 0.92 to 0.98 when it becomes the front stage. The design
mass flow will therefore be 64.2*0.98/0.92 = 68.4 kg/s.
The design pressure ratio of stages 3 to 10 inclusive will be given by the
figures tabulated above as:

Po*/P3 = (Po«/Pi)/(P3/Pi) = 7.665/1.696 = 4.52

If it was desired to operate stages 3 to 10 at the original design speed of


6487rpm it would be desirable to modify the casing so that the annulus area
ratios were the same as for the original design i.e. stage 10 outlet area would
have the same ratio to stage 3 inlet area as stage 7 outlet area has to stage 1 inlet
for the original design. The pressure ratio of stages 3 to 10 would then be the
same as stages 1 to 7 of the original design, that is 4.732 instead of 4.52. This
would require a modified casing and modified blade spans which may be
considered expensive. However at 6487 rpm the speed is approximately 9%
greater than the modified design speed of 5955 rpm and the performance at this
degree of overspeed may be acceptable if the somewhat greater mass flow and
pressure ratio are required, even if efficiency is likely to be a few points lower. If
further stages were to be removed from the front the degree of aerodynamic
overspeed when operating at 6487 rpm would too great to be tolerated without
restoring the stage matching by adjustment of the annulus areas.

Additional front stages

It is also possible to add stages to the front of the compressor to increase mass
flow and pressure ratio. Because the first rotor tip Mach number of the original
design is just subsonic it would be desirable not to increase this by more than is
essential, and so a constant outer diameter would be preferable for an additional
front stage, or ‘0’ stage as it is sometimes called. The hub diameter would
therefore rise through the ‘0 ’ stage, a feature which would help to alleviate the
somewhat higher blade loadings which would be required at the lower hub
diameter. It is possible that these increased loadings might lead to a decision to
Design for an industrial compressor 239

reduce the *0 * stage temperature rise, but assuming this is not deemed necessary
the design performance with this additional stage would be calculated as follows:

For original first stage at inlet MVT/P = 100*V288/101.3 = 16.75

Allowing for a drop of blockage factor when operating as the second stage:

M>/T/P = 16.75*0.95/0.98 = 16.24

To operate at this flow function and maintain the design stage matching the
original stages must operate at their design Mach numbers which means N/VT
must be as for the original design at the original first stage inlet i.e. 6487/^288 =
382.2. The temperature rises of all stages will increase somewhat due to their
increased inlet temperatures (AT/T will be as the original design not AT). When
two stages were removed in the previous example the speed was dropped by
about 8%; it is likely therefore that one additional front stage will require the
speed to be increased by approximately 4%, and therefore the stage temperature
will be increased by 8% since AT is proportional to N2. This gives a ‘0’ stage
design temperature rise of 29° compared to the original design value of 26.9°.
Hence the temperature at *0’ stage exit will be 317 K and the rotational speed
required will be 382.2*V317 = 6805. Now the 40f stage temperature rise will be
proportional to N2 so a second approximation to the temperature rise will be:

AT = 26.9(6805/6487)2 = 29.6°C

Hence T2 = 288+29.6 = 317.6 K and N*. = 6487(317.6/288),/2 * 6812

So AT = 26.9(6812/6487)2 = 29.7° and N ^ = 6813

For the ‘0* stage AT/T = 29.7/288 and T ^ i = 1.103

which for T| = 88% gives a stage pressure ratio of 1.353

At ‘O’ stage inlet MVT/P = (M>/Tj/P2)(P^P,XT,/T:)1'3

= 16.24*1.353/Vl.103 = 20.92

Hence: M = 20.92* 101.3/^288 = 124.8

Thus the mass flow is increased by almost 25% by the addition of the *0* stage.
However the rotational speed has been increased by 5% and this would increase
mechanical stresses by 10%, which would require to be allowed for. The
pressure ratio of stages 0 to 9 inclusive would be 8.915 compared to 7.665 for
240 Axialflow fans and compressors

the original 10 stages. Stages 0 to 10 inclusive ( 11 stages) would give a pressure


ratio of 10.37.
If mechanical considerations prohibited the 5% overspeeding, operation at the
original 6487 rpm would be satisfactory. Mass flow would be reduced by 6% to
7% due to the reduction of Va/U which inevitably occurs below design speed.
The pressure ratio of stages 0 to 9 inclusive would be approximately the same as
the original 10 stage compressor.

Scaled versions

For larger variations of the mass flow than readily obtained by the removal or
addition of stages increased or reduced scale versions of the compressor could be
used. The linear dimensions will be varied by the square root of the mass flow
change, i.e. for twice the design mass flow the dimensions will be increased by
V2 and the rotational speed will be reduced by the same factor so that the blade
speed is the same as for the original design. Some small adjustment to the mass
flow for the change of Reynolds number may be required. For a scaled version to
run at the synchronous speed of 3000 rpm (50 cycles per second) the linear scale
factor would be 6487/3000 = 2.162, which would give a first stage dp diameter
of 1.104*2.162 = 2.387 m. This is a large machine with a design mass flow of
100*2.1622 = 467 kg/s and a first stage blade of 0.477 m span and 0.238 m
chord. For a synchronous speed of 3600 rpm (60 cycles per second) the first sage
dp diameter would be 1.989 m, the blade span 0.398 m and the chord 0.198 m.
Some industrial manufacturers offer a wide range of scaled machines with the
largest diameter over three times the smallest; the mass flow of the largest being
of the order of ten times the smallest.

Volume flow

Rather than mass flow, industrial compressors are frequently specified by


volume flow.

From continuity: M = pAV, and so the volume flow, Q = M/p = AV

If A is the annulus area at entry to the first rotor, then V is the mass average axial
velocity. At the aerodynamic design speed (N/VT) and pressure ratio the value of
Va/U will be determined and so the volume flow will be proportional to the
mechanical speed irrespective of the inlet temperature and pressure. This makes
the volume flow a convenient measure of capacity for many purposes.
29 Outline design for a jet
engine compressor

Introduction

The basic requirements of a simple jet engine are small size and low weight for
the required thrust. In compressor aerodynamic terms this leads to a low inlet
diameter ratio and the fewest number of stages for the specified pressure ratio.
Whereas the simple jet engine was used for all types of jet propelled aircraft in
the early days of jet propulsion, today its use is confined to such applications as
unmanned aircraft and missiles where specific fuel consumption is less important
than low cost and size. The pressure ratio is low so that complicated surge
control devices are not required, probably not more than 5:1, and a mass flow in
the region of 10 kg/s in view of the modest thrust requirements.
So far as the compressor design is concerned the requirements are not radically
different from those for the fan of a low bypass ratio turbofan engine such as
used for a variety of modem military aircraft, except that the mass flow would
generally be much larger and the fan pressure ratio unlikely to exceed 3:1.

Specification

Mass flow = lOkg/s at T, = 288 K and P, = 101.3 kPa.


Pressure ratio = 4.5
Overall isentropic efficiency = 83% (polytropic efficiency = 86%)
Inlet diameter ratio = 0.5
Constant outer diameter through the stages.
Zero whirl velocity at all stator exits.

The specified efficiency is modest to allow for high blade inlet Mach numbers
in order to utilise high blade speeds and axial velocities to minimise the overall
diameter and number of stages. The constant outer diameter is specified for the
same reasons. The diameter ratio is not the minimum that could be specified but
242 Axialflow fans and compressors

allows for the possible addition of a *0 * stage if a significant thrust growth of the
basic design were required in the lifetime of the project. It also helps to avoid
excessive hub loadings while achieving a high first stage pressure ratio. Zero
whirl at stator exits is specified in order to avoid the need for an inlet guide vane
and is maintained through all stages in order to avoid excessive deflection in the
outlet stator.

Overall parameters

Number of stages

The overall temperature rise is given by:

AT = (Ti/n)(R(Y‘,)/y- 1) * (288/0.85X4.5° 286 - 1)

= 186.3

Reference to figure 3.9 indicates that AH/Va2 should not be greater than unity
and assuming for the moment equal temperature rises on each stage gives the
values of axial velocity in table 29.1.

Table 29.1

No. of stages 3 4 5
AT/stage °C 62.1 46.6 37.3
Vam/s 249.8 216.4 193.5
Inlet axial Mach No 0.777 0.664 0.588

Experience suggests an inlet axial Mach number of 0.65 is the highest practical
value. Thus 5 stages with an average temperature rise of 37.3° are chosen. The
temperature rise is reduced below the average on the first and last stage as these
stages suffer a greater variation of incidence than the middle stages. The first
stage also has higher loadings at the hub on account of the lower radius and
blade speed. Stage temperature rises are therefore provisionally chosen as in
table 29.2
Table 29.2

Stage 1 2 3 4 5
AT 34 38 40 39 36

It would be desirable to reduce the axial velocity through the stages in order to
shorten the diffuser between the compressor and combustor, however this would
require an extra stage if AH/Va2 is not to exceed unity.
Outline design for a jet engine compressor 243

Inlet annulus dimensions

A somewhat conservative value of 0.6 is chosen for the axial Mach number at
the first rotor inlet as it will meet the requirement for AH/Va2 not greater than
unity at the stage 1 temperature rise of 34°C.

For MN= 0.6, Va/VT = 11.6 and hence Va = 197m/s

MVT/AP = 34 and MVT/P = 10*V288 / 101.3 = 1.675

Hence Ap = 1.675/34 = 0.0493 m2 and allowing for boundary layers at inlet by a


blockage factor of 0.98 results in Ao = 0.0503 m2. The resulting diameters are:

Outer diameter, D = 292 mm


Inner diameter, d = 146 mm

First stage

Inlet Vai = 197 m/s, hence ti = 268.7 K and sonic velocity, a = 328.6 m/s
If the rotor tip inlet relative Mach No. is limited to 1.2, then:

Vj = 1.2*328.6 = 394.3 m/s

and Ua,*(394.3J 1972),/2 = 341.5 m/s

AHAJ’u, = 1005*34/341.5J = 0.293

and Va/Uy, = 0.577

Let stage 1 stator exit velocity, Va4 = 200 m/s to allow for the increasing
temperature rise of stages 2 and 3.

Stage 1 stator exit total temperature, T4 = 288+34 = 322 K

Hence Va/VT = 200/V322 = 11.15 and, from tables, MVT/AP = 33.1

Stage pressure ratio P4/P0 = OVTo) ^ 0 = (322/288)301 = 1.399

Hence stator exit MVT/P = 10*V322 / (1.399*101.3) = 1.266

So the stator exit flow area = 1.266/33.1 = 0.0382 m2 and taking a blockage
factor of 0.95 gives Ao = 0.0403 m2 and a hub diameter of 184 mm.
244 Axialflow fans and compressors

At the rotor exit, assuming the mid-span diameter * 0.8*292 = 233.6 mm and
that Va = 197m/s, the stator inlet velocity is:

V, = (Va1 + v W ) m and Vw, = AH/U since Vw0 = 0

which gives: V3 = 237.4 m/s and lj = T3 - (V32/2C,) = 294 K

The density ratio across the rotor is given to a close approximation by:

P2/P 1 = (t 2/t,)lw lr,)l''-(294/268.7)! = 1.197

From continuity, A2/A 1= piVaj/pjVai = 0.835

and A2 = 0.0493*0.835 = 0.0412 m2

Because the majority of the static pressure rise occurs in the rotor due to the
relatively high reaction of a zero inlet whirl stage the same blockage factor is
assumed as at stator exit and the geometric area is therefore 0.0412/0.95 =
0.0434 m2 which results in a hub diameter of 173 mm and a blade speed of
341.5*173/292 = 202.3 m/s at the rotor trailing edge.

Rotor hub

The rotor hub inlet and outlet vector diagrams will be as in figure 29.1. As has
been indicated in Chapter 7 there will be an increase of relative total temperature
through the rotor hub due to the increasing radius towards the trailing edge.

AT«, = AH«i/Cp = (U22 - U,:)/2CP= (202.32 - 170.82)/(2*1005) = 5.8°C

Figure 29.1 Rotor hub vector diagrams

When the blade loading is considered in terms of the static pressure recovery
factor, Cpi, the pressure equivalent of this temperature rise may be deducted, i.e.
Outline design for a jet engine compressor 245

the total and static temperatures at rotor exit are both assumed to be 5.8° lower
than their actual values. Thus from the vector diagram of figure 29.1:

t2 = t3 = To +34 - S tflC f = 288 + 34 - 33.5 = 288.5 K

The effective static temperature t‘2 = 288.5 - 5.8 = 282.7 K

Because the rise of relative total temperature is removed:

T, = T, = T0 + (V,J - Va,l)/2Cp = 302.5 K

Cp, = (P2 - p,)/(Pi - P.) = [(ti/T,)3S- (ti/T,)35l / [1 - <t,/T,)35] = 0.378

If a value for Cpi is calculated without allowing for the rise of relative total
temperature a value of 0.55 is obtained, which might be considered unacceptably
high, whereas the effective value of 0.378 is conservative.

Last stage

Space does not permit the detailed design of all five stages to be examined so the
last stage will now be considered. The axial velocity at the fourth stator exit can
be slightly reduced to 190 m/s in view of the fifth stage temperature rise of 36°C.
The total temperature rise to this station is 150° and the local total temperature
and pressure are therefore calculated as:

Stage 5 entry total temperature = 288+150 = 438 K

The total pressure ratio to this station will be:

P,/P, = (438/288)’ 01 =3.53

and P3= 3.53* 101.3 = 357.6 kPa

MVT/P = 10*^438/357.6 = 0.585

Va/VT= 190/^438 = 9.08

and so, from tables, MVT/AP = 28.5

Flow area, Af = 0.585/28.5 = 0.0205 and using a blockage factor of 0.88 gives:

Ao = 0.0205/0.88 = 0.02333 mJ
246 Axialflow fans and compressors

From which the inner diameter is calculated as 236mm and the diameter ratio is
0.807. The axial velocity will be assumed constant across the fifth rotor as an
initial assumption, and reduced if this proves desirable.

The temperature ratio across the stage = 1 + 36/438 = 1.0821

The pressure ratio = 1.08213'01 = 1.268 for a stage efficiency of 86%

Assuming a mean radius equal to 0.9 of the tip radius gives:

U = 0.9*341.5 = 307.4 m/s

Since Vw0 = 0; Vw3= AH/U = 1005*36/307.4 = 117.7 m/s

t, = ti = T j-V j2/2C,

= 474 - (1902+ 117.7J)/(2* 1005)

= 449.1 K

t, = to = T0 - Va2/2Cp

= 438 - 190I/(2*I005) = 420K

Since Va is taken as constant across the rotor

A,p/A2f = P2/Pi = (tjA,)2 = 449.1/420 = 1.0692

Ajf = 0.0205/1.0692 = 0.0192 m2

and Am = 0.0192/0.88 = 0.0218 m2

The inner diameter is therefore 240 mm at rotor exit and 236 mm at entry which
results in a increase of relative total temperature of only 0.95°C at the rotor hub.
This compares to a figure of 5.8°C across the first rotor hub.

The stator inlet angle at mid-span, (X3 = tan’^Vwj/Va)

= tan*,(l 17.7/190) = 31.7°

Deflecting the flow to the axial through the final stator at constant axial
velocity would therefore represent a low static pressure recovery factor and it
would be advantageous to drop the axial velocity at outlet so far as reasonable to
Outline design for a jet engine compressor 247

shorten the following diffuser. Let the de Haller number for the mid-span flow
be 0.75, which represents a modest diffusion. The stator inlet velocity is:

Vj = (190* + 117.7l)'n = 223.5 m/s

dH = 0.75 = V</Vj

Hence: V, = 0.75*223.5 = 168 m/s

V«/\/T= 168/^474 = 7.72

and,: MVT/AP = 24.92 from tables

MVT/P = 10*^474/(4.5* 101.3) = 0.478

Ar = 0.478/24.92 = 0.01918 m1

and Ao “ 0.01918/0.88 = 0.218

This give a hub diameter of 240 mm at the stator oudet, assuming the outer
diameter is maintained constant at 292 mm. By coincidence this means a
constant hub diameter through the stage 5 stator but had the figure been
marginally different a constant hub diameter would have suggested itself.

Blading

The blade geometry would be determined in a similar manner to the preceding


examples of Chapters 27 and 28. Rotor blading could have DCA profiles
throughout and stators could be DCA, C4 or NACA 65. If the necessary
computer codes were available all stators and rotors of stage 3, 4 and 5 could
have controlled diffusion aerofoils with some performance advantage. Rotors for
the first two stages, where tip Mach numbers are supersonic would remain DCA
profiles.
30 HP compressor mid-stage

Introduction

In multistage aeroengine compressors, when the flow is fully developed after


three or four stages, it a common practice in the blade design to take account of
the spanwise variation of the flow conditions through the annulus boundary
layers. One system for doing this was discussed in Chapter 14 and this will now
be applied to a stage of a high pressure compressor.

Datum stage

It is convenient to consider the design as a modification of a datum stage


designed for constant work and constant efficiency along the span. A mid stage
of 0.85 diameter ratio is considered where the local design conditions are as
listed in Table 30.1.

Table 30.1

Inlet total pressure = 2000kPa


Inlet total temperature = 785 K
Cp = 1059 J/kgK
Y= 1.372
Zero whirl at stator exit along the span.
Blade tip speed = 388 m/s
Axial velocity = 200 m/s constant along the span.
AT = 46.27°C constant along the span,
rj = 88% constant along the span.
HP compressor mid-stage 249

Radius ratio r/r,

Figure 30.1 Spanwise efficiency variation

Blade end modifications


The blade design is modified to take account of the increased total pressure
losses towards the ends of the span. For this purpose a variation of the stage
efficiency according to figure 30.1 is employed and the stage work adjusted to
produce a constant increment of total pressure. At the same time the axial
velocity is reduced towards the blade ends and increased at mid span while
maintaining substantially the same inlet relative velocities to the blades as for the
datum design.

AH*, = Cp AT = 1059*46.27= 49005 r a V

I1<hi AHi* = 0.88*49005 = 43124 m V

AHmod = ndnAHju Almod s 43124/r| mod


Using the efficiencies of figure 30.1 gives the necessary work variation along
the span to satisfy the requirements of the modified design. Considering the tip
and hub radii:

AHmod = 43124/0.7 = 61606 m2/s2


It is appropriate to maintain the same reaction as for the datum design at each
radius and therefore the increased AH should be equally disposed either side of
the mid point between the datum triangle vertices as illustrated in figure 30.2
When the reaction is 50%, as in the diagrams of figure 14.5, the axial velocities
before and after the rotor remain equal when the work is increased at constant
inlet relative velocities to the rotor and stator. In the present case, where the
250 Axialflow fans and compressors

A*W2U

Datum
Modified

AfU,/2U
Figure 30.2 Datum and modified blade end vector diagrams

reaction is greater than 50%, this is not the case and the inlet axial velocity to the
rotor falls more than the inlet axial velocity to the stator, if both inlet relative
velocities are held constant. There is no necessity, however, for the relative inlet
velocities to be kept precisely constant, provided that the pressure rise factor in
the blades is not unduly increased. Since, in the present case, the pressure rise
factor for the rotor will be the greater the inlet relative velocity to the rotor will
be held constant and that at inlet to the stator allowed to fall slightly to give
equal axial velocities before and after the rotor. This also has the advantage of a
somewhat greater fall of axial velocity from the datum which will favour the
mass integrated efficiency.
The whirl velocity leaving the upstream stator hub is given by:

Vwo = (AIW2U) - (AHmod^U) = (49000 - 61600)/(2*329.8) = -19.1 m/s


This velocity is negative because it is against the direction of the blade speed.
Vlrd = VlfBl ^ = (2002 + 329.82)i/2 = 386 m/s
Vwlmod = U - Vw0= 329.8 - (-19.1) = 348.9 m/s
V a ^ = 3862 - 348.91 = 165.1 m/s

mod = tan (V\V| mod/VOfDod) = 64.7°

(Xomod = tan [(U/Vamod) - tancti mod] = ■6.7

mod= tan [tancti mod - (AHmod/UVamod)] —44.6

a 3 mod = tan'1[(U/Vam0d) - tana2m0d] = 45.3°


HP compressor mid-stage 251

V3mod = Va^od /cosa3mod= 234.7 m/s

Similar calculations can be made for other radii, noting that the work is
reduced from the datum at the mid radius and the axial velocity increased. At
intermediate radii of 88.5% and 96.5% of the tip radius the vector diagrams will
be unchanged as the efficiency equals the datum value of 88% at these radii. A
comparison of the datum and modified air angles is shown in figure 303.

Radius ratio r/rt

Figure 303 Datum and modified air angles

Rotor design

The hypothesis discussed in Chapter 7 for blade design where secondary losses
are high is employed. Considering the rotor hub section first, the relevant angles
are:
a, =64.8° a 2 = 44.6°

The ideal static pressure recovery factor is:

Cpi = 1 - (cos<X|/cos<X2)2 = 0.642

The stage total pressure loss is the sum of rotor and stator losses, hence:

(APr +APs)/p = (1 - H)AH = (1 - 0.7)61600 = 18480 m2/s2

This loss must be shared between the rotor and stator and a reasonable
approximation is to assume that the loss coefficients are the same.

Thus: APrM p Vi2 = APsMpV32 and since (Vs/V,)2 = 0.37 we obtain:


252 Axialflow fans and compressors

APnMpV,* = APjMpVj2= 18482/(1.37*0.5*385.5J) = 0.1815

Hence Cp*, = 0.642 -0.1815 = 0.46

It is reasonable to assume that the minimum profile loss coefficients of the


rotor and stator are equal with a value of 0.025. Hence:

C p * =0.46 + 0.025 = 0.485

The effective inlet angle is given by:

cos2ai eff = cos2a 2(l - Cpidr) = cos244.6(1 - 0.485) = 0.261

From which we obtain <Xi * = 59.3°. The blade design is therefore carried out for
aie<r= 59.3° and eta = 44.6°. Using the methods of McKenzie(1980):

tana* = (tan59.3 + tan 44.6)/2 = 1.335

Since a middle stage of a high pressure compressor requires only a modest stall
margin the stagger angle can be chosen for maximum efficiency at the design
point
Hence: tan? = tana*, - 0.213, from which £ = 48.3°

Since the effective inlet angle is 5.5° less than the actual inlet angle it will be
preferable to design for a small negative value of incidence to the effective inlet
angle in order to keep the actual incidence less than +5°. After some trials -1° is
found to be suitable, giving a camber of 24° at a S/C of 0.918.
Similar calculations for the tip section suggest a camber of 24.7° at S/C of
0.981, corresponding to an effective incidence of -1° and the actual incidence is
-4.5°. The ratio of true tip chord to hub chord is therefore 1.1, which is
satisfactory, and the ratio of tip axial chord to hub axial chord is 0.833, which is
also satisfactory. A linear variation of true chord with radius gives:

S/C = 0.561 +0.42r/r«ip

For the mid span section it is assumed there are only profile losses and so the
effective angles are the actual angles. At radii of 0.885 and 0.965 of the tip
radius, where the vector triangles are unchanged, it might be assumed that the
blading would be unaltered from the datum since the efficiency is the same for
the modified and datum designs. However this is not stricdy correct as the losses
are somewhat greater than the profile losses alone. The calculations are therefore
made as for the hub and tip sections for a stage efficiency of 88%.
HP compressor mid-stage 253

Radius ratio r/r,

Figure 30.4 Rotor geometry for constant and variable efficiency with radius

Comparison with conventional blade geometry is most readily appreciated if the


conventional design is calculated for the same space/chord ratios as the modified
blading. The results are shown on figure 30.4 in terms of stagger and camber
angles.

Rotor aspect ratio

The equivalent cone angle of the mid span blade geometry is given by:

tanoteq = [(l/70(h/c)(S/C)],/2[cosct2,/2 - cosct|,/2)

For S/C = 0.95, a, = 59.6°, a 2 = 47.5°, and a*, = 4.0°; h/C = 1.32, which would
be a suitable value. The mid-span diffusion factor is 0.4 and plotting on figure
6.6 for an aspect ratio of 1.32 and clearance to chord of 0.02 to 0.04 indicates a
satisfactory stall margin and a moderately low end wall loss parameter.

Stator blade

The stator blade geometry is calculated by the same methods but because of the
low oudet angles, despite the mid stage requirement for a modest stall margin,
the stagger angles for blades of such low outlet angle are preferably obtained
from tan£ = tana™ -0.15 rather than using the constant 0.213. For the hub
section this results in the following possible range of blade geometry:

a 3 = 45.6°; cu = -6.6°; and APu«MpV32mod = 0.1815 (as for the rotor)

hence Cp, = 0.504 and Cp, eff = 0.504 - 0.1815 + 0.025 = 0.348
254 Axialflow fans and compressors

and so: a 3etT= 36.6°, and £ = 9.3°

Table 30.1

-5 0 +5
fc 41.6 36.6 31.6
e 64.6 54.6 44.6
s /c 0.47 0.254 N/A
Wr +4 +9 N/A

Thus a somewhat high camber of 64.6° must be accepted to avoid too low a
S/C and also keep the effective and actual incidences within the range -5° to +5°.
For stators it is convenient to keep the axial chord constant along the span, if this
gives acceptable blade sections at all radii. The stator geometry at other radii can
be calculated on this basis as in the table below.

Table 30.2

T/Tch 0.85 0.885 0.925 0.965 1.0


<*3 45.6 35.5 32.5 33.2 40.9
<*3eff 36.6 33.7 32.5 31.6 33.4
cu -6.6 0 +1.3 0 -5.7
C 9.3 10.4 10.2 9.0 7.4
s/c 0.47 0.487 0.51 0.534 0.555
e 64.7 44.0 39.0 39.6 56.7
fr 41.6 32.4 29.6 28.8 35.7
Icff -5.0 +1.3 +2.9 +2.8 -2.3
1 .4 +4.0 +3.1 +2.9 +4.4 +5.2

With the exceptions of the high camber at the hub section and a marginally
high actual incidence at the casing, these figures are satisfactory and the blade
design is acceptable. Figure 30.5 shows curves for the stagger, camber and S/C
of the datum and modified stator blades. Comparison with figure 30.4 shows that
the stagger changes at the blade ends are in opposite directions for rotor and
stator, rotor staggers are increased and stator staggers reduced by the
modifications.

Stator aspect ratio

Based on the mid span blading the equivalent cone angle is only 2° at an aspect
ratio of one and 3.4° at an aspect ratio of three. The choice of aspect ratio might
therefore be based on other considerations such as natural frequency of vibration
or blade numbers. A value of about 1.5 would be anticipated and thickness to
HP compressor mid-stage 255

chord ratios of 8% at the casing reducing to 6% at the inner diameter would be


typical.

Radius ratio r/r,

Figure 30.5 Stator camber, stagger and space/chord

Design method

Although the above design method for variable spanwise efficiency has not been
directly proven experimentally, some designs with similar geometrical features
are known to have shown improved performance. Although it is easy to ascribe a
performance change to the design system used, it must be borne in mind that it is
the resulting blade geometry that determines the performance; if two different
design systems produce similar blades then the performance will be similar
despite differences in the assumed vector geometry or blade selection data.
31 The high bypass fan

Introduction

It is not practical to carry out a modem design exercise for this type of fan
without the use of axisymmetric throughflow and blade to blade computer
programs. A description of a typical design has been made available by courtesy
of Rolls Royce pic and this gives an insight to the form such designs take and the
significant differences between them and other types of axial flow fans and
multistage compressor blading. Two aerodynamic features are most obvious; the
pressure ratio varies markedly along the span, and the tip relative Mach number
is about 1.5.

General arrangement

The outer part of the blades of this type of fan provide directly 70% or more of
the total propulsive thrust of a modem turbofan engine designed for subsonic
transport aircraft. The flow through the hub section supercharges the core
compressors and typically is 10% to 25% of the total flow, giving a bypass ratio
(defined as the ratio of bypass to core mass flow) in the range from 3 to 9. For
optimum propulsive efficiency the pressure ratio required of the fan increases as
the bypass ratio diminishes and below a ratio of 3 the optimum pressure ratio is
greater than can be provided by a single stage fan. Above a bypass ratio of 8 the
installed drag of the fan tends to mitigate the further apparent advantage of
increased propulsive efficiency. The result is the almost universal adoption of a
single stage fan for modem subsonic transport turbofan aeroengines.
As has been discussed previously the absence of inlet guide vanes is a
significant advantage in weight, cost and complexity for such a fan. Because the
driving turbine is of appreciably smaller tip diameter than the fan it is desirable
to use a high blade speed for the fan in order to minimise the number of turbine
stages required to drive it. The combination of these factors led to the
The high bypass fan 257

development of the transonic single stage fan with a corrected dp speed in the
region of 450 m/s and a relative inlet tip Mach number of approximately 1.5. The
diameter ratio at the fan inlet may be as low as 0.3 in order to achieve the highest
practical mass flow per unit frontal area. Immediately following the rotor the
flow is split into bypass and core streams, each stream having its own stator
blade row. The result is an arrangement as shown in figure 31.1 where it should
be noted that the large axial gap between the rotor and the stators helps to reduce
noise due to interaction of the blade rows.

Figure 31.1 Annulus and blading arrangement of high bypass fan

Aerodynamic parameters

Some of the leading aerodynamic parameters are listed in table 31.1 for three,
stream lines through the rotor and it should be noted that the outlet radius is
some 20% greater than the inlet radius at the hub, and 2% less at outlet
compared with inlet at the tip. The hub value of AH/U2 is 1.276 which is high,
but the associated de Haller number is alleviated by the increase of relative total
temperature and pressure due to the increase of radius through the rotor.
The inlet axial Mach numbers are high to give a high value of flow per unit
area in order to minimise the total frontal area. Despite the lack of any whirl
velocity ahead of the rotor there is a small variation of static pressure radially
(shown by values of p/Po, where Po is the freestream inlet total pressure) due to
meridional curvature. The slightly reduced value of P/Po near the tip is due to an
allowance for the casing boundary layer. The loss coefficients near hub and tip
258 Axialflow fans and compressors

make allowance for the concentration of secondary losses in these regions and
shock losses at the tip.

Table 31.1
High bypass fan aerodynamic parameters

Rotor inlet
Po = 37 kPa; To = 259.6 K

SstifiD_______________ HsacJmk________ MisLssao________ Near tip

Radius radorM is 0.327 0.665 0.973


Blade speed m/s 143 292.5 428.3
Axial vel. m/s 200 231 184
Axial Mach No. 0.641 0.755 0.589
Relative Mach No. 0.798 1.21 1.493
p/P0 0.76 0.686 0.77
P/P0 1 1 0.982

Across the rotor


Total loss coeff. 0.147 0.053 0.2
S/C 0.32 0.594 0.77
Diff. factor 0.474 0.588 0.52
Vao*/Vata 0.809 0.687 0.767
Stream height ratio 0.874 0.877 0.864
W rj. 1.206 1.04 0.978
AH/U2** 1.276 0.569 0.4
Po«/Pto 1.542 1.842 1.955
Rotor efTy 0.894 0.945 0.785

Rotor exit
Radius ratio r/rtia 0.394 0.691 0.952
Blade speed m/s 173.5 304.2 418.9
Axial velocity m/s 161 158.8 141.2
Absolute Mach No. 0.846 0.695 0.625
p/Po 0.969 1.31 1.464

Air angles
ax 35.9 51 66.8
02 -17.6 39.5 60.6
ct3 54 47.4 49.9
The high bypass fan 259

The values of S/C are low and this allows the diffusion factors to be kept
below the maximum recommendation of 0.6 while also ensuring, near the tip,
sufficient suction surface length behind the shock to provide for subsonic
boundary layer pressure rise, as discussed in Chapter 9. Near the hub the S/C is
lower than even the high deflection of 53.5° may require and this is to give an
acceptable ratio of tip chord to hub chord for stress purposes.
The axial velocity falls significantly across the rotor, the ratio being as low as
0.687 at mid-span, however, the stream tube height ratio is 0.877 due to the
density ratio. Along the stream tube the mass flow is constant so we have:

M = piA|Vai = p2A2Va2 and A = 2nrh

Hence: Vai = r« 5h|/r25h2

where 5h is the radial height of a small streamtube. The LHS of this last equation
is known as the axial velocity density ratio or AVDR. Where r2.= r1 this reduces
to the streamtube height ratio. Except near the hub, the ratio 5h2/5h| is a close
approximation to the AVDR in the present case.
The total pressure ratio, PMt/Pia. rises from only 1.5 at the hub to nearly 2 at the
tip. This is due primarily to the low hub blade speed limiting the pressure ratio,
but it is also the case that any reduction of pressure ratio at the tip would tend to
reduce the efficiency due to the need to have a high static pressure recovery
factor to achieve a high efficiency. This is apparent from the diffusion efficiency
expression:
Tldiff=l -0/Cpi

where G5 = - pi) and Cpi = Apa/(P|- Pi)

This indicates that unless the loss coefficient is reduced in proportion to the ideal
pressure recovery factor the efficiency will diminish. It is desirable therefore to
maintain the pressure recovery factor at as high a level as practical and this must
result in a high pressure ratio when the inlet Mach number (and therefore the
dynamic pressure) is high.
Under 'Rotor exit* it should be noted that the radius is quoted as a ratio to the
inlet tip radius which allows direct comparison of the streamline radius at blade
inlet and outlet. While there may be some small variation of the absolute Mach
number between rotor trailing edge and the stator leading edge it is clear that the
stator inlet Mach numbers will be subsonic. The values for p/P0 indicate a very
significant increase of static pressure from hub to casing as is to be expected for
equilibrium of the high whirl velocities generated by the rotor.
The air angles listed indicate a significant negative rotor exit relative angle,
near the hub, consistent with the higher than unity value for the work coefficient.
The values of a 3 will vary somewhat between rotor trailing edge and stator inlet.
260 Axialflow fans and compressors

it is clear, however, that to obtain zero whirl at the stator exits will require large
deflections at all radii.
It would appear from some of the fan rotor blade aerodynamic parameters that
these fans can operate to higher loadings than downstream blade rows. This may
owe something to the fact that the inlet flow to the fan is completely free of
wakes from upstream blades or struts which, for other blade rows, cause a rapid
fluctuation of the incidence.

Rotor blade geometry

The transonic fan rotor blade has probably been the single item of greatest axial
compressor aerodynamic research and development effort over the past 25 to 30
years. This is justified by the fact that typically a 1% increase of the bypass
stream efficiency gives a 0.7% reduction of specific fuel consumption which, for
a large aircraft, is a saving in the region of £70,000 per aircraft per year. The
bypass stage efficiency at a pressure ratio of 1.75 has been progressively raised
from the low 80’s to a little over 90%. Most of this has come from the fan rotor
and some of the steps in this achievement are:

• Reduced S/C and thinner aerofoils


• Full streamline analysis
• Wide chord without mid-span shroud
• Multi arc profiles replacing DCA
• Inflected camber line for reduced shock loss
• Full three dimensional flow computation

Multi arc profile has low


camber over first 50% of
chord and higher camber
over second 50%.

Inflected profile gives


improved shock pattern
with reduced losses at
maximum Mach No.

Figure 31.2 Supersonic fan blade profile development

Fan blade integrity

Throughout these aerodynamic developments the requirements for blade


integrity had to be preserved both for steady stress, vibration and foreign object
The high bypass fan 261

ingestion. The latter particularly is opposed to the aerodynamic preference for


thin blades with sharp leading edges.
Most early designs featured an aspect ratio in excess of three and required a
mid-span shroud to prevent stalled vibration at part speed. This feature itself led
to a complex coupled flap and torsional vibration at high speeds and much
research and development was necessary to ensure this vibration did not occur
below the maximum corrected rotational speed encountered in flight. The
introduction of wider chord blades, with aspect ratios less than 2.5, with hollow
construction at the larger sizes (above about 1.5 m tip diameter) allowed the
removal of the mid-span shroud, obviating the high speed vibration mode and
improving the efficiency. This also reduced typical numbers of rotor blades from
33 to 22.
On the experimental side the aerodynamic research was enhanced by the use of
laser anemometry, which provided a non-intrusive method of measuring the
velocity patterns within the blade passages. Comparison of these measurements
with computed passage velocity patterns allowed rapid improvement of the
computations, thus reducing the need for much costly trial and error
development.
Bibliography

Adkins Q G, Smith L H. (1982), *Spanwise mixing in axial flow turbomachines'


Trans. ASME Journal of Engineering for Power Vol. 104 pp 97-110.

Adkins R C, Jacobsen O H, Chevalier P. (1983), 4A preliminary study of annular


diffusers with constant diameter outer walls*, ASME paper No. 83-GT-218.

Andrews S J. (1949), Tests related to the effect of profile shape and camber line
on compressor cascade performance'. Aeronautical Research Council R and M
2743.

Andrews S J, Jeffs R A, Hartley E L. (1956), Tests Concerning Novel Designs


of Blading for Axial Compressors', Aeronautical Research Council R A M 2929
HMSO.

Azimian A R, McKenzie A B, Elder R L. (1989), ‘Application of Recess Vaned


Casing Treatment to Axial Flow Fans. ASME paper No. 89-GT-68.

Bammert K, Staude R. (1979), ‘Optimization for Rotor Blades of Tandem


Design for Axial Flow Compressors’, ASME Paper No.79-GT-125.

Bullock R O. (1964), ‘Analysis of Reynolds Number and Scale Effects on


Performance of Turbomachinery’, ASME Journal of Engineering for Power, July
1964 pp 247-256.

Camp T R, Horlock J H. (1993), ‘An Analytical Model of Axial Compressor


Off-Design Performance’, ASME Paper No. 93-GT-96.

Cargill A M, Freeman C. (1990), ‘High Speed Compressor Surge with


Application to Active Control’, ASME paper 90-GT-354.
Bibliography 263

Carter A D S , Moss C E, Green G R, Annear G C. (1957), ‘The Effect of


Reynolds Number on the Performance of a Single Stage Compressor', National
Gas Turbine Establishment Report No. R 204.

Cheshire L J. (1945), ‘The design and development of centrifugal compressors


for aircraft gas turbines', Proc. I. Mech. E. Vol. 153.

Cohen H, Rogers G F C, Saravanamuttoo H I H. (1972), Gas Turbine Theory,


Longman.

Cumpsty N A. (1989), Compressor Aerodynamics. Longman Scientific and


Technical.

Day I J. (1976), Axial Compressor Stall, University of Cambridge PhD


dissertation.

Day I J, Greitzer E M, Cumpsty N A. (1977), 'Prediction of Compressor


Performance in Rotating Stair, ASME paper 77-GT-10.

Day 1 J. (1991), 'Active Suppression of Rotating Stall and Surge in Axial


Compressors), ASME paper 91-GT-87.

Denton J D. (1978), 'Throughflow calculations for transonic axial flow


turbines’. ASME Journal of Engineering for Power Vol.100 pp. 212-18.

Dimmock N A. (1961), ‘A Compressor Routine Test Code’, National Gas


Turbine Establishment Report NoJl246.

Dring R P, Spear D A. (1990), ‘The Effects of Wake Mixing on Compressor


Aerodynamics’, ASME paper 90-GT-132.

Emmons H W, Pearson C E, Grant H P. (1955), ‘Compressor Surge and Stall


Propagation’, Trans ASME Vol.27 pp. 455-469.

Elder R L, Gill M. (1984), ‘A Surge Prediction Method for Multistage Axial


Flow Compressors’, I.Mech.E. conference on Computational Methods in
Turbomachinery, University of Birmingham. Paper C74 April 1984.

Ffowcs Williams J E, Harper M F L, Allwright D J. (1992), ‘Active Stabilisation


of Compressor Surge in a Working Engine’ ASME paper 92-GT-88.

Freeman C, Dawson R E. (1983), ‘Core compressor development for large civil


jet engines’, Paper 83 Tokyo IGTC-46 International gas turbine congress Tokyo
October 1983.
264 Axialflow fans and compressors

Freeman C. (1985), ‘Effect of tip clearance flow on compressor stability and


engine performance’, von Karman Institute for Fluid Dynamics. Lecture Series
1985-05.

Gallimore S J, Cumpsty N A. (1986), ‘Spanwise mixing in multistage axial flow


compressors Part I Experimental investigation*, Trans. ASME Journal of
Turbomachinery vol. 108 pp. 2-9.

Gallimore S J. (1986), ‘Spanwise mixing in multistage axial flow compressors


Part: II Throughflow calculations including mixing*, Trans ASME Journal of
Turbomachinery Vol 108 pp. 10-16.

Greitzer E M. (1978), ‘Surge and Rotating Stall in Axial Flow Compressors Part
II Experimental results and comparison with theory*, ASME Journal of
Engineering for Power Vol 98 April 1978 pp. 199-217.

Gresh M T. (1991), Compressor Performance. Selection, Operation and Testing


of Axial and Centrifugal Compressors. Butterworth-Heinemann.

Hill R J, Nicholas D J, Tubbs H. (1990), ‘Compressing the Compressor*, Dr.D


M Smith memorial lecture. LMechE. March 1990.

Hobbs D E, Weingold H D. (1983), ‘Development of Controlled Diffusion


Airfoils for Multistage Compressor Application* ASME Paper 83-GT-211.

Horlock J H. (1978), Actuator Disc Theory. McGraw-Hill

Howell A R, (1944), ‘A note on Compressor base aerofoils*. Power Jets


Memorandum No. M 1011.

Howell A R. (1945), ‘Fluid Dynamics of Axial Compressors* Proc.lMech£.


Vol. 153 pp. 441-482.

Howell A R, Bonham R P. (1951), ‘Overall and Stage Characteristics of Axial-


flow Compressors*, Proc.LMechE.

Howell A R, Calvert W J. (1978), ‘A New Stage Stacking Technique for Axial-


Row Compressor Performance Prediction*, ASME Paper No. 78-GT-139.

Howell W T. (1964), ‘Stability of Multistage Axial Flow Compressors’,


Aeronautical Quarterly XV Nov. 1964.

Huppert M C, Benser W A. (1953), ‘Some stall and surge phenomena in axial


flow compressors’, Journal of Aeronautical Sciences, 1953 Vol. 20 No.12.
Bibliography 265

Ivanov S K, Dudkin V E, Molchanov V N. (1984), ‘Axial flow ventilation fan’,


UK Patent Application No. 2 124 303 A. Filed 4 July 1983, Published 15
February 1984.

Lakshminarayana B, Suryavamshi N, Prato J, Moritz R. (1994), ‘Experimental


Investigation of the Flow Field in a Multistage Axial Flow Compressor', ASME
paper No. 94-GT-455.

Lieblein S.(1965),. ‘Experimental Flow in Two-Dimensional Cascades’, Chapter


VI in The Aerodynamic Design of Axial Flow Compressors NASA report SP 36.

Livesey J L, Hugh T. (1966), ‘Suitable Mean Values in One-dimensional Gas


Dynamics', Journal of Mechanical Engineering Science Vol.8 No.4.

McKenzie A B. (1963), ‘Multi-Stage Axial Flow Compressor (Interstage bleed


off-take)*, UK Patent Specification 936,635.

McKenzie A B. (1980), ‘The design of axial compressor blading based on tests


of a low speed compressor*, Proc. I. Mech. E. Vol.194 pp 103-111.

McKenzie A B. (1984), ‘Encounters with Surge*, von Karman Institute for Fluid
Dynamics. Lecture series 1983-84 Unsteadyflows in turbomachines.

McKenzie A B. (1988), ‘The selection of fan blade geometry for optimum


efficiency', Proc. I. Mech. E. Vol.202 Al.

McKenzie A B. (1993), ‘Passive stall control for fans and compressors’,


IMech£. seminar on turbo compressor andfan stability, 8 April 1993.

Miller D C. (1977), ‘The Performance Prediction of Scaled Axial Compressors


from Model Tests’, I.Mech.E. 1977.

Miller D C, Wasdell D L. (1987), ‘Off-design prediction of compressor blade


losses' l.Mech.E paper C279I87.

Moyes S J. (1956), ‘A Method of Correction of Compressor Characteristics for


Change of Working Fluid Specific Heat', National Gas Turbine Establishment
Memorandum No.276.

Nicholas D J, Freeman C. (1982), ‘Recent Advances in the Performance of High


Bypass Ratio Fans’ 13th ICAS Congress Tokyo, August 1982.

Moody L F. (1944), ‘Friction factors for Pipe Flow', Trans. ASME Vol.66 1944
p.671.
266 Axialflow fans and compressors

Palmer J, Ramsden K, Goodger E. (1987), Compressible Flow Tables for


Engineers. Macmillan Education Ltd.

Pearson H, McKenzie A B. (1959),'Wakes in Axial Compressors', Technical


Notes, Journal of the Royal Aeronautical Society July 1959.

Reid C. (1969), 'The Response of Axial Flow Compressors to Intake Flow


Distortion*, ASME Paper 69-GT-29.

Robbins W H, Dugan J F. (1965), 'Prediction of Off-Design Performance of


Multistage Compressors', Chapter X in The Aerodynamic Design of Axial Flow
Compressors' NASA report SP 36.

Robinson C J. (1985), ‘Measurement of the Vortex Hows in a Low Speed Axial


Flow Compressor’, AJAAS5-1360.

Ruffles P C. (1991), 'Keys to Excellence in Design and Manufacture’, Lecture at


the inauguration of the University Technology Centre in Gas Turbine
Combustion, Cranfield Institute of Technology 30 July 1991.

Schaffier A. (1979), 'Experimental and Analytical Investigation of the Effects of


Reynolds Number and Blade Surface Roughness on Multistage Axial
Compressors', ASME paper 79-GT-2.

Schweitzer J K, Garbeloglio J E. (1984), 'Maximum loading capability of axial


flow compressors’, AIAA Journal of Aircraft Vol.21 pp 593-600.

Shepherd D G. (1956), Principles of Turbomachinery. The MacMillan


Company, New York.

Smith G D J, Cumpsty N A. (1984), 'Flow Phenomenon in compressor casing


treatment’, ASME Journal of Engineering for Gas Turbines and Power Vol.106
pp. 532-41.

Smith L H. (1969), 'Casing boundary layers in multistage compressors’, Flow


research on blading. Ed. Dzung L S. Elsevier Publishing Company.

Stenning A H. (1973), 'Rotating Stall and Surge’, Fluid Dynamics of


Turbomachines. ASME lecture course, Iowa State University.

Tubbs H, Rae A J. (1991), 'Aerodynamic development of the high pressure


compressor for the IAE V2500 aero engine’, I. Mech. E. European Conference
19-20 March 1991 Turbomachinery - Latest developments in a changing scene.
Paper C423/023.
Bibliography 267

Wassel A B. (1967), ‘Reynolds Number Effects in Axial Compressors* ASME


Paper No. 67 - WAJGT-2.

Wennerstrom A J. (1986), ‘Low Aspect Ratio Axial Flow Compressors: Why


and What it Means’, Third Cliff Garrett Turbomachinery Award Lecture SAE
SP-683.

Wright P I, Miller D C. (1991), ‘An improved compressor performance


prediction model*. I.Mech.E. European Conference 19-20 March 1991
Turbomachinery - Latest developments in a changing scene. Paper No.
C423/028.

Ziabasharhagh M. (1992), ‘Recess Vane Passive Stall Control for Axial Flow
Fans*, PhD thesis Cranfield University 1992.

Ziabasharhagh M, McKenzie A B, Elder R L. (1992), ‘Recess Vane Passive Stall


Control*, ASME paper No. 92-GT-36.

You might also like