282
, - -----~ AEDC-TR-77-80 ARCHIVE COPY- DO NOT LOAN .f AN ANALYSISOF THE INFLUENCE OF SOME EXTERNALDISTURBANCES ON THE AERODYNAMIC STABILITY OF TURBINEENGINEAXIAL FLOW FANS AND COMPRESSORS m m=D mo- ==:m~ m B __ "n Z E~ = E ~ ENGINE TEST F/~CILITY ARNOLD ENGINEERING DEVELOPMENT CENTER AIR FORCE SYSTEMS COMMAND ARNOLD AIR FORCE STATION, TENNESSEE 37389 August 1977 Final Report for Period June 1973 -- June 1977 Approved for public release; distribution unlimited, ,'.', :" J; CI'Ce Prepared for DIRECTORATE OF TEST ENGINEERING DEPUTY FOR OPERATIONS ARNOLD ENGINEERING DEVELOPMENT CENTER AIR FORCE SYSTEMS COMMAND ARNOLD AIR FORCE STATION, TENNESSEE 37389 .o.. .'

t, I ARCHIVE COPY- DO NOT LOAN

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

I

t, t -

.t

. - - - - - ~

AEDC-TR-77-80 ARCHIVE COPY- DO NOT LOAN

. f

AN ANALYSIS OF THE INFLUENCE OF SOME

EXTERNAL DISTURBANCES ON THE AERODYNAMIC

STABILITY OF TURBINE ENGINE AXIAL FLOW

FANS AND COMPRESSORS

P - - m

m = D m o - ==:m~

m B ~___ "n

" - " Z

E ~ =

E ~

ENGINE TEST F/~CILITY ARNOLD ENGINEERING DEVELOPMENT CENTER

AIR FORCE SYSTEMS COMMAND ARNOLD AIR FORCE STATION, TENNESSEE 37389

August 1977

Final Report for Period June 1973 -- June 1977

Approved for public release; distribution unlimited,

,'.', :" J; C I ' C e

Prepared for

DIRECTORATE OF TEST ENGINEERING DEPUTY FOR OPERATIONS ARNOLD ENGINEERING DEVELOPMENT CENTER AIR FORCE SYSTEMS COMMAND ARNOLD AIR FORCE STATION, TENNESSEE 37389

i . . . o . . . '

,- • .,.,: -:

NOTICES

When U. S. Government drawings specifications, or other data are used for any purpose other than a definitely related Government procurement operation, the Government thereby incurs no responsibility nor any obligation whatsoever, and the fact that the Government may have formulated, furnished, or in any way supplied the said drawings, specifications, or other data, is not to be regarded by implication or otherwise, or in any manner licensing the holder or any other person or corporation, or conveying any rights or permission to manufacture, use, or sell any patented invention that may in any way be related thereto.

Qualified users -may obtain copies of this report from the Defense Documentation Center.

References to named commercia! products in this report are not to be considered in any sense as an endorsement of the product by the United States Air Force or the Government.

This report has been reviewed by the Information Office (OI) and is releasable to the National Technical Information Service (NTIS). At NTIS. it will be available to the general public, including foreign nations.

APPROVAL STATEMENT

Tl~s technical report has been reviewed and is approved for publication.

FOR THE COMMANDER

MARION L. LASTER Director of Test Engineering Deputy for Operations

AEAN L. 7~E~VEREAU~"

Colonel, USAF Deputy for Operations

• -=' . ; : . . . . . . . :'~'.,~r..': . . . . . ~,."

UNCLASSIFIED REPORT DOCUMENTATION PAGE

t R E P O R T NUMBER J~ GOVT ACCESSION NO,

I AEDC-TR-77-80

4 TqTLE ( ~ d Subrrlte)

AN ANALYSIS OF THE INFLUENCE OF SOME EXTERNAL DISTURBANCES ON THE AERODYNAMIC STABILITY OF TURBINE ENGINE AXIAL FLOW FANS AND COMPRESSORS ? A u T H O R f s J

William F. Klmzey, ARO, I n c .

9 P E R F O R M I N G ORGAN~ZAT |ON NAME AND ADDRESS

A r n o l d E n g i n e e r i n g D e v e l o p m e n t C e n t e r A i r F o r c e S y s t e m s Command A r n o l d A i r F o r c e S t a t i o n , T e n n e s s e e 37389 I L C O N T R O L L I N G O F F I C E NAME AND ADDRESS

A r n o l d E n g i n e e r i n g D e v e l o p m e n t C e n t e r (XRFIS) A r n o l d A i r F o r c e S t a t i o n , T e n n e s s e e 37389 14 M O N ' T O R t N G AGENCY NAME & ADDRESSr | I d l l l e ren r / tom C o n t r o l l | n 8 0 ( h c e )

READ iNSTRUCTIONS BEFORE COMPLETING FORM

5 R E C I P I E N Y ' S C A T A L O G NUMBER

S T Y P E 0 F R E P O R T & P E R I O D C O V E R E D

F i n a l R e p o r t - J u n e 1973 - J u n e 1977

6 P E R F O R M I N G ORG R E P O R T NUMBER

B C O N T R A C T OR G R A N T NUMBER(e )

10 PROGRAM E L E M E N T P R O J E C T , TASK AREA & WORK UNIT NUMBERS

P r o g r a m E l e m e n t 6 5 8 0 7 F

1 :) R E P O R T D A T E

A u g u s t 1977 13 NUMBER O F PAGES

271 IS S E C U R I T Y CLASS. (e l IhJe report)

UNCLASSIFIED

ISe DECL ASSI F I C A T I O N DOWNGRADING SCNEOuLE N/A

16 D I S T R I B U T I O N S T A T E M E N T (o f th le Repor t )

A p p r o v e d f o r p u b l i c r e l e a s e ; d i s t r i b u t i o n u n l i m i t e d .

17 D I S T R I B U T I O N S T A T E M E N T (o f fhe n b l t r a c t enlered In B lock 30, I I d i f fe ren t .

l i t S U P P L E M E N T A R Y N O T F..~

A v a i l a b l e Ln DDC

19 K E Y WORO$ tCon tmue on reve r i e side f l n e c e s s e ~ ~ d i d e n t l l y by bloch number)

m a t h e m a t i c a l a n a l y s i s e n g i n e s c o m p r e s s o r s e x t e r n a l t u r b i n e d i s t u r b a n c e s a x i a l f l o w a e r o d y n a m i c s t a b i l i t y f a n s

2G J B S " R A C T (Cont inue ~ r e v e r i e side I ! necessary ~ d Jden|mfy ~ bJock numbe~

The o v e r a l l o b j e c t i v e o f t h e work d e s c r i b e d h e r e i n was t o d e v e l o p an i m p r o v e d m e t h o d f o r t h e c o m p u t a t i o n o f t h e i n f l u e n c e o f e x t e r n a l d i s t u r b a n c e s on t u r b i n e e n g i n e c o m p r e s s o r s t a b i l i t y , a n d , i n s o d o i n g , t o g a i n a d d e d p h y s i c a l i n s i g h t i n t o t h e d e s t a b i l i z i n g p r o c e s s e s . The o b j e c t i v e s w e r e a c c o m p l i s h e d t h r o u g h t h e d e v e l o p m e n t o f a o n e - d i m e n s i o n a l , t i m e - d e p e n d e n t m a t h e m a t i c a l c o m p r e s s o r m o d e l f o r a n a l y s i s o f p l a n a r d i s t u r b a n c e s and an

FORM 1473 E D I T I O N O F I N O V 6 S I S O B S O L E T E DD , JAN 73

UNCLASSIFIED

UNCLASSIFIED

20. ABSTRACT ( C o n t i n u e d )

e x t e n s i o n o f t h e model t o a t h r e e - d i m e n s i o n a l fo rm f o r a n a l y s i s o f d i s t o r t e d i n f l o w s . The m o d e l s s a t i s f y m a s s , momentum and e n e r g y e q u a t i o n s on a t i m e - d e p e n d e n t b a s i s . C o m p r e s s o r s t a g e f o r c e and s h a f t work w e r e d e t e r m i n e d f rom e m p i r i c a l s t a g e c h a r a c t e r i s t i c s w i t h c o r r e c t i o n s made f o r u n s t e a d y c a s c a d e a i r f o i l a e r o d y n a m i c s . The s y s t e m o f e q u a t i o n s c o m p r i s i n g t h e mode l i s s o l v e d u s i n g a d i g i t a l c o m p u t e r . Example p r o b l e m s w i t h c o m p a r i s o n s t o e x p e r i m e n t s a r e p r e s e n t e d f o r t h r e e d i f f e r e n t c o m p r e s s o r s . Example p r o b l e m s s o l v e d u s i n g t h e o n e - d i m e n s i o n a l a n a l y s i s i n c l u d e : d e t e r m i n a t i o n o f t h e s t e a d y - s t a t e s t a b i l i t y l i m i t ( s u r g e l i n e ) w i t h u n d i s t u r b e d f l o w , i n s t a b i l i t y c a u s e d by o s c i l l a t i n g p l a n a r i n f l o w , dynamic r e s p o n s e o f a c o m p r e s s o r t o o s c i l l a t i n g e n t r y p r e s s u r e , dynamic r e s p o n s e t o o s c i l l a t i n g d i s c h a r g e p r e s s u r e , and c o m p r e s s o r i n s t a b i l i t y c a u s e d by r a p i d upward ramps o f e n t r y t e m p e r a t u r e . Example p r o b l e m s s o l v e d u s i n g t h e d i s t o r t i o n model i n c l u d e : s t a b i l i t y l i m i t ( s u r g e ) l i n e r e d u c t i o n c a u s e d by a combined r a d i a l and c i r c u m f e r e n t i a l p r e s s u r e d i s t o r t i o n , t i m e - v a r i a n t d i s t o r t i o n e f f e c t s , p u r e r a d i a l p r e s s u r e d i s t o r t i o n e f f e c t s , and p u r e p r e s s u r e and t e m p e r a t u r e c i r c u m f e r e n t i a l d i s t o r t i o n e f f e c t s . C o m p a r i s o n s a r e made w i t h e x p e r i m e n t a l r e s u l t s , and t h e m o d e l s d e v e l o p e d a r e shown t o c o m p u t e t h e c o m p r e s s o r s t a b i l i t y l i m i t s w i t h r e a s o n a b l e a c c u r a c y .

A F S r A , ~ l r AFIL T e n q

UNCLASSIFIED

A E D C-T R -77-80

PREFACE

The research reported herein was partially performed

by the Arnold Engineering Development Center (AEDC), Air

Force Systems Co,and (AFSC). The work and analysis for

this research were done by personnel of ARO, Inc., AEDC

Division (a Sverdrup Corporation Company), operating

contractor for the AEDC, AFSC, Arnold Air Force Station,

Tennessee, under Program Element 65807F. A portion of the

work was conducted under ARO Projects No. RF435-12YA,

R32P-98A, R32P-12YA, R32P-50A, and R32P-C5A. The author

of the report is William F. Kimzey, ARO, Inc. Eules L.

Hively is the Air Force project manager. The manuscript

(ARO Control No. ARO-ETF-TR-77-47) was submitted for

publication on July 12, 1977.

The author wishes to acknowledge Mr. Fran Loper, ARO,

Inc., for his assistance in the development of numerical

analysis and computer programs required for this effort;

Mr. Don Couch and Mr. Ted Garretson, ARO, Inc., for their

work in computer program development and in running some

of the data presented in this report; and Dr. B. H. Goethert,

University of Tennessee Space Institute, for assistance

provided during the course of this work. The material re-

ported herein was submitted to the University of Tennessee as

partial fulfillment of the requirements of the degree of

Doctor of Philosophy.

A EDC-TR-7?-80

CONTENTS

Chapter Pag___~e

I. INTRODUCTION . . . . . . . . . . . . . . . . . . 1

Ramifications of Stability to Engine

Operation . . . . . . . . . . . . . . . . . . 3

Major Operating Conditions and External

Disturbances Influencing Stability ..... 3

Previous Analyses of Compressor Stability

with Undisturbed Flow . . . . . . . . . . . . 8

Previous Analyses of Compressor Stability

with Disturbed Flow . . . . . . . . . . . . . 12

Objectives and Approach of Current Work .... 19

Organization . . . . . . . . . . . . . . . . . 22

II. DEVELOPMENT OF ONE-DIMENSIONAL,

TIME-DEPENDENT MODEL . . . . . . . . . . . . . 23

Governing Equations . . . . . . . . . . . . . . 24

Stage Forces and Shaft Work . . . . . . . . . . 29

Method of Solution . . . . . . . . . . . . . . 43

III. EXAMPLE PROBLEMS WITH COMPARISON TO

EXPERIMENTS, ONE-DIMENSIONAL MODEL ...... 60

Steady-State Stability Limits . . . . . . . . . 61

Dynamic Stability Limits . . . . . . . . . . . 66

IV. DEVELOPMENT OF THREE-DIMENSIONAL MODEL

FOR DISTORTED FLOWS . . . . . . . . . . . . . . 78

Governing Equations . . . . . . . . . . . . . . 79

iii

AEDC-TR-77-80

Chapter

Simplifications . . . . . . . . . . . . . . . .

Special Considerations for Distortion

Effects . . . . . . . . . . . . . . . . . . .

Method of Solution . . . . . . . . . . . . . .

V. EXAMPLE PROBLEMS WITH COMPARISON TO

EXPERIMENTS, DISTORTION MODEL . . . . . . . . .

Combined Distortion Pattern,

XC-1 Compressor . . . . . . . . . . . . . . .

Time-Variant Distortion . . . . . . . . . . . .

Pure Radial and Circumferential

Distortion Patterns . . . . . . . . . . . . .

VIo SUMMARY OF RESULTS AND RECOMMENDATIONS

FOR FURTHER WORK . . . . . . . . . . . . . . .

One-Dimensional, Time-Dependent Analysis . . .

Distortion Analysis . . . . . . . . . . . . . .

Recommendations for Further Work . . . . . . .

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . .

Page

86

92

i01

109

109

115

117

122

123

126

129

133

A E DC-TR-77-80

ILLUSTRATIONS

Figure Page

1. Typical Aircraft Gas Turbine Compression

System Configurations . . . . . . . . . . . . . 142

2. Typical Compressor Map . . . . . . . . . . . . . 143

3. Effect of Flight Altitude and Mach Number on

Compressor Operating Region . . . . . . . . . . 144

4. Typical Influences of External Disturbances

on Compressor Stability . . . . . . . . . . . . 146

5. Surge Analysis Based on Compressor and

Throttle Curve Shapes . . . . . . . . . . . . . 149

6. Representative Axial Flow Compressor

Arrangement and Operation . . . . . . . . . . . 150

7. Physical System Modeled and Control Volumes. . . 151

8. Rotor Fluid Element Velocities and Forces. . . 152

9. Representative Stage Characteristics [46] .... 153

I0. Steady-State and Unsteady Stage

Characteristic Behavior . . . . . . . . . . . . 154

ii. Unsteady Airfoil and Cascade Aerodynamics .... 155

12. Comparison of Theoretical and Experimental

Unsteady Cascade Lift Coefficients ...... 156

13. Variation of Dynamic Lift Coefficient Ratio

with Mean Angle of Attack . . . . . . . . . . . 157

14. Unsteady Lift Coefficient Magnitude Ratio .... 158

Y

AEDC-TR-77-80

ri~e

15. Computational Plane Designations for Finite

Difference Solutions . . . . . . . . . . . . . 159

16. Solution Procedure for One-Dimensional,

Time-Dependent Model . . . . . . . . . . . . . 160

17. Station Locations (Computation Planes) for

One-Dimensional, Time-Dependent Models .... 161

18. XC-I Compressor Model Loaded to Stall,

N//~ = 90% . . . . . . . . . . . . . . . . . . 162

19. Flow Breakdown Process Leading to Instability,

XC-I Compressor, N//~ = 90% .......... 163

20. Comparison of Model Computed 90% Speed

Characteristic and Stability Limit with

Experimenual Measurements, XC-I Compressor . . 164

21. NACA-8 Compressor Model Loaded to Stall,

N//~ = 90% . . . . . . . . . . . . . . . . . . 165

22. Comparison of Model Computed Speed Character-

istics and Stability Limits with Experimental

Measurements, NACA-8 Compressor ........ 168

23. Comparison of Model Computed Flow Coefficients

at Compressor Stability Limit with

Experimental Measurements,

NACA-8 Compressor . . . . . . . . . . . . . . . 169

24. Comparison of Model Computed Speed Character-

istics and Stability Limits with Experimental

Measurements, J85-13 Compressor ........ 171

vi

AE DC-T R-77-80

Figure Page

25. Oscillating Inflow Loading to Stall,

NACA-8 Compressor, N//~ = 70% . . . . . . . . . 172

26. Stability Limits for Planar Oscillating

Inflow, NACA-8 Compressor . . . . . . . . . . . 179

27. Test Arrangements for NASA J85-13 Dynamic

Response Experiments . . . . . . . . . . . . . 180

28. Variation of Amplitude Ratio and Phase Angle

Through the J85-13 Compressor with

Oscillating Compressor Entry Pressure

at 40 Hz . . . . . . . . . . . . . . . . . . . 181

29. Dynamic Response to Oscillating Compressor

Entry Pressure, J85-13 Compressor ....... 183

30. Variation of Amplitude Ratio and Phase Angle

Through the J85-13 Compressor with

Oscillating Engine Fuel Flow at 40 Hz . . . . 189

31. Dynamic Response to Oscillating Fuel Flow,

J85-13 Compressor . . . . . . . . . . . . . . . 191

32. Rapid Inlet Temperature Ramp to Stall,

NACA-8 Compressor . . . . . . . . . . . . . . . 193

33. Compressor Instability Caused by Rapid Inlet

Temperature Ramps . . . . . . . . . . . . . . . 196

34. Division of Compressor into Control Volumes. . . 197

35. Control Volume Velocities and Fluxes ...... 198

36. Circumferential Crossflow . . . . . . . . . . . . 199

37. Radial Crossflow . . . . . . . . . . . . . . . . 200

vii

AE DC-TR-77-80

Figure Page

38. Radial Distortion Effects . . . . . . . . . . . . 201

39. Circumferential Distortion Effects ....... 202

40. Compressor Entry Total Pressure Distribution

for Combined Distortion Pattern ........ 205

41. Compressor Entry Pressure Input Values for

Combined Distortion Pattern . . . . . . . . . . 206

42. Comparison of Computed and Experimental

Stability Limits with Combined Distortion

Pattern, XC-I Compressor . . . . . . . . . . . 207

43. Circumferential Profiles in Tip Region with

Combined Distortion Pattern,

XC-I Compressor 208 I m a l o o D e e m e e i i o m

44. Comparison of Computed and Measured

Circumferential Distortion Attenuation for

Combined Pattern . . . . . . . . . . . . . . . 210

45. Radial Profiles with Combined Distortion

Pattern . . . . . . . . . . . . . . . . . . . . 213

46. Radial Profiles with Uniform Entry Flow ..... 215

47. Flow Coefficient Distribution Just Prior to

Instability with Combined Distortion

Pattern . . . . . . . . . . . . . . . . . . . . 217

48. Time-Variant Distortion . . . . . . . . . . . . . 221

49. Relative Frequency Spectrum for Time-Variant

Distortion Pattern . . . . . . . . . . . . . . 222

50. Loss in Stability Limit Pressure Ratio with

Time-Variant Distortion . . . . . . . . . . . . 223

°,°

ym

AEDC-TR-77-80

Figure Page

51. Computed Radial Distortion Influences on

Stability . . . . . . . . . . . . . . . . . . . 224

52. Comparison of Radial Distortion Model

Computations with Experimental Results .... 226

53. Model Computed Influence of Circumferential

Pressure Distortion on Stability,

XC-I Compressor . . . . . . . . . . . . . . . 227

54. Comparison of Circumferential Distortion

Model Computations with Experimental

Results . . . . . . . . . . . . . . . . . . . . 228

55. Circumferential Crossflow Magnitudes and

Effects . . . . . . . . . . . . . . . . . . . . 229

56. Model Computed Influence of Circumferential

Distortion Angular Extent on Stability,

XC-I Compressor . . . . . . . . . . . . . . . . 231

57. Effect of Unsteady Airfoil Stall Properties

on Stability Margin . . . . . . . . . . . . . . 232

58. Model Computed influence of Circumferential

Temperature Distortion on Stability,

XC-I Compressor . . . . . . . . . . . . . . . . 233

59. Comparison of Circumferential Temperature

Distortion Model Computations with

Experimental Results . . . . . . . . . . . . . 234

AED C-TR-77-80

APPENDIXES

A. FIGURES FOR CHAPTERS I THROUGH VI . . . . . . . .

B. COMPRESSOR DETAILS AND STAGE

CHARACTERISTICS . . . . . . . . . . . . . . . .

C. FIGURES AND TABLES FOR APPENDIX B . . . . . . . .

NOMENCLATURE . . . . . . . . . . . . . . . . . . . . .

141

235

237

265

X

A E DC-T R-77-80

CHAPTER I

INTRODUCTION

Stable aerodynamic operation of the compression

system of an aircraft gas turbine engine is essential for

the engine to operate safely and to satisfactorily deliver

the performance which it is designed to provide. The com-

pression system of an aircraft gas turbine engine consists

of one or more compressors arranged in configurations such

as those illustrated in Fig. 1 (Appendix A). 1 Aerody-

namically stable operation refers to the condition in which

the compressor delivers a desired quantity of airflow on a

continuous basis, compressed to a desired pressure level,

free of excessively large amplitude fluctuations in the flow

properties throughout the compressor. The aerodynamic

stability limit is most commonly represented as a locus of

corrected rotor speed points denoting the minimum corrected

flow rate value for stable operation on a plot of the total

pressure ratio across the compressor versus corrected air-

flow rate as shown in Fig. 2. Operation to the right of the

curve is stable and to the left of the curve is unstable.

The stability limit curve is often referred to as

the surge line or stall line of the compressor stemming from

IAII figures in Chapters I through VI appear in Appendix A.

A E DC-T R-77-80

the fact that surge is one possible specific form of

instability experienced in compressors, and all compressor

instabilities are initiated by aerodynamic stall of some

blades or vanes of the compressor.

The stability margin of the compressor is commonly

taken to be the percentage difference between the compressor

total pressure ratio at the stability limit and that at the

operating line holding corrected airflow constant (Fig. 2).

The position of the operating line is primarily determined

by the load imposed on the compressor by downstream engine

components. However, under actual flight conditions in an

aircraft, the positions of both the stability limit line and

the operating line are influenced by external disturbances

in the flow at either the entry or exit of the compressor.

The overall objective of this study is to develop

an improved analytical method which will allow prediction of

the loss of stability margin of the compressor caused by

external disturbances imposed upon it by flight conditions

in an aircraft. Also, the development of the method is

intended to provide some degree of physical understanding of

the processes which lead to compressor instability.

Prior to beginning the analysis it is helpful to

review the ramifications of stability on engine operation

and to review some of the experimentally observed influences

of the operating conditions and external disturbances on

compressor stability. Also, a review of some of the

AEDC-TR -77-80

previously accomplished works directed at stability pre-

dictions both with and without external disturbances will be

made.

I. RAMIFICATIONS OF STABILITY TO ENGINE OPERATION

Operation of the compressor to the left of the

stability limit (Fig. 2) will result in loss of thrust,

possible loss of engine control, or possible engine

structural damage caused by overheating and high cyclic

stress induced in the engine by unstable flow conditions.

Reference [1] 2 presents evidence of thermal and cyclic

stress effects imposed by compressor instability for one

particular engine. Montgomery [2] has compiled a history of

difficulties in the operation of several aircraft gas

turbine engines which have been experienced because of

compressor stability problems. Montgomery's work well

illustrates the need for a method of determining the sta-

bility limit of the gas turbine engine compressor,

especially as they are affected by the conditions of flight.

II. MAJOR OPERATING CONDITIONS AND EXTERNAL

DISTURBANCES INFLUENCING STABILITY

The position of the stability limit line and the

operating line of the compressor of a given design are

2Numbers in brackets refer to similarly numbere~ references in the Bibliography.

AE DC-T R-77-80

affected in flight by the condition of the flow entering the

compressor and by the throttling or back pressure provided

by the compressor or engine component downstream of the

compressor under consideration. The entry total tempera-

ture, the Reynolds number at which the compressor must

operate, and the nonuniformity and unsteadiness of the flow

at the compressor inlet have major influences on the sta-

bility of the compressor.

Because of the variation of total temperature at

the compressor entry which occurs with flight Mach number

and altitude, the compressor's operating region moves far

from the design point of the compressor; thus, a relatively

large region of the compressor map must be considered in any

analysis of stability. Figure 3a depicts the variation of

operating region with compressor inlet total temperature for

a simple single spool turbojet compressor. Figure 3b shows

the influence of flight Mach number and altitude on engine

entry total temperature. The changes in flight conditions

cause the Reynolds number to change. Figure 3c gives an

example of the Reynolds number influence on the stability

limit. The Reynolds Number Index (REI) used in Figure 3c is

the ratio of the Reynolds number in flight to that which

would exist at sea-level-static conditions at the same

compressor inlet axial Mach number. Figure 3d shows the

variation of REI with flight condition.

Figure 4 shows experimentally determined losses in

stability margin caused by nonuniform (distorted) flow

4

AE DC-TR-77-80

conditions and unsteady flow conditions at the compressor

entry plane. In Fig. 4 the loss is illustrated by downward

shifts of the stability limit toward the operating line. In

some cases, the disturbances also cause stability margin

loss by moving the operating line upward. For purposes of

illustration, the entire loss is shown in Fig. 4 as a down-

ward movement of the stability limit. Figure 4a from [4] is

for the case of steady but distorted flow into the com-

pressor created by a wire screen of nonuniform porosity

placed in front of the compressor. Figure 4b from [5]

illustrates the loss in stability margin caused by uniform

but sinusoidally time-varying flow into the compressor

produced by a rotating flow interrupter upstream of the

compressor. The case of combined unsteady and distorted

flow, sometimes called time-variant distortion, is shown in

Figs. 4c and 4d for a turbojet engine compressor [6] and a

turbofan engine fan [7]. The engine entry flow fields were

produced by a venturi-centerbody arrangement which produced

an unstable shock wave boundary layer interaction which

causes the adverse flow field at the engine entry plane [6].

Figure 43, based on data from [8], indicates the influence

on the downstream compressor stability limit line of the

nonuniform and unsteady flow field exiting from a forward

compressor of a turbofan engine.

In addition to disturbances in the pressure enter-

ing the compressor, total temperature disturbances may occur.

Information from [9] shown in Fig. 4f, indicates the

A EDC-TR-77-80

experimentally determined influence of time-steady spatial

variation o[ total temperature entering the compressor.

Rapid uniform total temperature increases at the engine

inlet which could be caused by ingestion of hot gases pro-

duced by gun or rocket fire can also cause a loss of

stability [10].

Disturbances to the compressor may originate at

the exit of the compressor as well as at the inlet. An

experiment conducted on a turbofan engine described in [ii]

demonstrated the influence of an oscillating back pressure

on fan stability.

The distortion examples of Fig. 4 all use total

pressure or total temperature variation to describe the

distortion pattern characteristics and to correlate with

loss in stability margin. The use of total pressure and

total temperature has been shown through experience and is

accepted in practice to be the most practical distortion

pattern description method (see Bibliography). The

Society of Automotive Engineers Aerospace Recommended

Practice on Gas Turbine Engine Inlet Flow Distortion

Methodology [12] specifies total pressure as the parameter

for distortion pattern description. Total pressure and

total temperature are used because they are much more

readily measured than velocity distribution, static pressure

and flow direction. Also, when a distorted flow field

approaches an engine, the interaction of the engine with the

distortion pattern produces a specific static pressure,

6

AEOC-TR-77-80

velocity and flow angularity pattern at the engine entry

(di&cussed in Chapter IV). Thus, the combination of a total

pressure or total temperature distortion pattern with an

engine at a given operating condition implicitly specifies

a velocity, static pressure and flow direction distribution

at the compressor entry and throughout the compressor.

Measurement of velocity and flow direction at the compressor

entry is sometimes suggested as a more physically meaningful

set of measurements for distortion pattern definition

because rotor blade angle of attack is related to the

velocity vectors at the rotor blade leading edge. However,

compressor instability caused by distortion can (and often

does) occur because of stall in any stage of the compressor.

Therefore, unless detailed velocity and flow direction

measurements are made at each and every compressor stage

entry, no real additional information is provided by the

more difficult velocity and flow direction measurements.

Further, the total pressure and total temperature distri-

butions produced by an aircraft inlet, if measured approxi-

mately an engine diameter ahead of the compressor entry

plane, will be the same with or without an engine installed

behind the inlet. Therefore, distortion pattern information

from scale model and full-scale inlet tests may be obtaaned

without the engine installed. This could not be done if

velocity static pressure and flow direction were chosen as

the distortion pattern definition parameters. For these

reasons, total pressure and total temperature variations

AE DC-TR-77-80

will also be used for disturbance description for the

analytical work in this study.

Most of the influences shown in Fig. 4 were

determined from simulations of actual flight in altitude

test cells or wind tunnels using artificial methods of dis-

turbance generation. These disturbances are, however,

representative of actual disturbances experienced in flight.

The influence on stability in actual flight is, however,

more difficult to measure; thus, the ground test sources are

used. More detailed information on the generation of these

disturbances in flight is given in [13 through 16].

III. PREVIOUS ANALYSES OF COMPRESSOR STABILITY

WITH UNDISTURBED FLOW

Early analysis of compressor instability dealt

with "surging" described as "a violent fluctuation in

delivery" by Whittle in an early work (1931) on the turbo-

compressor as a supercharger [17]. In an even earlier work

by Kearton [18] and in later works by Den Hartog [19] and

IIorlock [20], analysis of surge was performed using the

shapes of the pressure-ratio-mass flow curves of the com-

pressor and the throttle (back pressuring device) to locate

stable and unstable combinations of the compressor and

throttle and to qualitatively describe surging. (Figure 5

is a simple illustration of this analysis.) With the

throttle set to an open condition, the match point, A, is a

stable operating condition. This may be deduced by

8

AE DC-T R-77-80

imagining that the flow is momentarily perturbed upward, if

this should happen, the upward perturbation will cause a

reduction in the pressure exiting from the fan, and an

increase in the resistance to flow from the throttle. Both

of these effects tend to cause flow to decrease, thus

countering the original perturbation and restoring equi-

librium. A similar argument applies for a downward

perturbation of flow rate. If, however, the match point is

point C, instability will occur. Once more, imagine an

upward perturbation of flow. This will result in an

increase in compressor discharge pressure which is greater

than the increase in resistance to flow; thus, the upward

perturbation in flow will be reinforced causing the flow to

continue to increase. Thus, point C is unstable. This

analysis may be generalized to state that the compressor-

throttle combination will be unstable in regions where the

slope of the compressor characteristic curve is greater than

the slope of the throttle loss characteristic curve.

Applying this rule, if operation is attempted near point C,

and the flow is perturbed upward, it will continue to

increase until it reaches point B where the flow increase is

countered (that is, the compressor characteristic slope

becomes less than the throttle curve slope) and, ignoring

dynamic effects, is caused to begin to decrease. Once the

flow starts down, however, it will continue to decrease

until it reaches a region near point D where the slope of

the compressor characteristic is once more less than that of

9

A EDC-TR-77-80

the loss characteristic. Thus, a limit cycle, quaiitatively

describing surge based on characteristic curve nonlinearity,

but ignoring dynamic effects, is constructed.

This simple explanation provided a reasonable

qualitative explanation of the surge instability, but in

actual use tended to predict a value of flow rate at which

surge would be encountered which was often lower than that

actually experienced [21]. It later became common practice

to assume that instability would occur at the peak com-

pressor ratio; that is, at the first point where the

compressor characteristic slope became positive. This

assumption was reasonably well borne out for high corrected

rotor speeds by experiments [21, 22].

While the theories described to this point were

based principally on observations made on centrifugal com-

pressors, they are in principle applicable to axial flow

units. (For example, the compressor characteristic curve of

Fig. 5 in no way specifies the type of compressor which

produced it.) Pearson and Bowmer (1949) [21] specifically

addressed the problem of stability in axial flow compressors.

Pearson and Bowmer observed the discrepancy which existed

between compressor experiments and the simple nonlinear

characteristic curve theory and noted that the nonlinear

theory ignored the dynamic effects caused by volumetric

capacitance, fluid column inertia, and flow resistance in

the compressor and ducting. Pearson and Bowmer approached

the problem by deriving an approximate set of dynamic

I0

AEDC-TR-77-80

(time-dependent) equations for a twelve-stage axial flow

compressor and related the compressor to an analogous

electrical circuit. The compressor was stacked from assumed

stage characteristics to form a representative multi-stage

unit. It was then assumed that instability would occur at

any condition where the impedance of any given part of the

circuit representing a stage was zero. For positive

impedance, disturbances in the circuit would be damped; for

negative impedance the disturbances would be amplified, thus

causing instability.

Howell (1964) [23] extended the work of Pearson

and Bowmer by applying the same technique of drawing an

electrical analogy to the compressor, but used a more

detailed stability analysis on the resulting electrical net-

work. Howell also applied the analysis to an actual

compressor and compared those results to the experimentally

determined stability limit. Howell's analysis produced a

reasonably good prediction of the stability limit, but it

was somewhat below the experimentally observed limit at

design rotor speed.

Kimzey and Couch (1969) developed a stability

criterion for predicting the position of the stability limit

directly from an analysis of a simple set of dynamic

equations representing the volumetric capacitance and

inertial characteristics of the compressor on a stage-by-

stage manner [24]. This stability criterion was based on

the assumption that the compressor stages work in pairs,

I!

AEDC-TR-77-80

and if any "pair" becomes an unstable combination, then

instability of the entire compressor will result. (Any

given stage in the compressor is a member of two pairs; one

is formed in conjunction with the preceding stage and one

with the following stage; for the last stage, the ducting

and throttle form one pair; for the first stage, the up-

stream ducting forms a pair with the first stage.) The

criterion was applied to an actual compressor and compared

to experimental data. Reasonably good agreement was shown.

Daniele, Blaha, and Seldner (1974) [25] applied

more sophisticated stability criteria to an approximate set

of linearized dynamic equations representing volumetric

capacitance, inertial, and energy effects. These criteria

were applied to a compressor for which detailed experimental

data were available to predict the steady-state stability

limit line. Very good agreement with the experimental data

was achieved.

IV. PREVIOUS ANALYSES OF COMPRESSOR STABILITY

WITH DISTURBED FLOW

The work described above was principally directed

at the prediction of the compressor stability limit with

steady and uniform flow entering the compressor. Work has

also been performed in an effort to predict the influence of

external disturbances on the stability limit. An excellent

and comprehensive review of the work accomplished to date in

this area is given in [16].

12

AE DC-T R-77-80

Stcad~ Nonuniform (Distorted) Flow

Notable early work dealing with the influence of

distortion at the compressor entry was accomplished by

Alford (1957) [26]. Alford treated the problem of circumfer-

ential total pressure distortion with a penetration theory

which postulated that the influence of circumferential

distortion on stability margin was proportional to the depth

of penetration of the distorted pattern into the compressor.

The distortion pattern attenuation characteristics were

computed in an approximate manner, thereby deriving the form i

of an index which correlated distortion pattern character-

istics to loss in stability margin. Alford concluded that

the loss in margin depended on the magnitude of the total

pressure deficit in the low pressure region, on the area of

the compressor entry annulus which had total pressure below

the average value, and on the angular extent of the low

pressure region.

Pearson (1963) [27] described the parallel com-

pressor theory of circumferential distortion. The theory

divides the compressor annulus into arc sectors, each sector

being viewed as an independent compressor running in

parallel with the other sectors. All sectors are assumed to

discharge to a common static pressure. The pressure ratio-

flow rate characteristic curve of each sector was assumed to

be unchanged by the distortion. Crossflows between the

adjacent parallel compressor sectors were ignored. Each

sector will have a different inlet total pressure because of

13

A E D C-T R-77-80

the circumferential distortion existing at the compressor

entry plane. In its simplest form of application, the

entire compressor is assumed to roach its stability limit

when the stability limit of any sector is encountered. In

practice, correlating indexes, such as described in [28] are

developed which relate the actual loss in stability margin

to the size of the arc sector chosen and the size of the

deficiency of total pressure in the lowest pressure sector.

Total temperature distortion has also been treated using

parallel compressor theory [28].

Goethert and Kimzey [29] presented an analog

computer model for calculating the effect of circumferential

distortion which was principally based on the parallel com-

pressor theory, but did make provisions for approximating

crossflow effects between sectors. Fett (1969) [24]

combined parallel compressor theory with the Hurwitz

stability criteria to provide a method of directly pre-

dicting the loss of stability margin caused by circumfer-

ential distortion for simple two- and three-stage

compressors.

Comparisons between the degree of distortion

predicted to cause instability by the parallel compressor

theory in its simplest form and that observed experimentally

indicated that as the angular extent of the low pressure

region was reduced from 180 degrees, the compressor would

actually become more tolerant to the distortion than pre-

dicted.

14

AE DC-T R-77-80

Reid (1969) [30], and Calogeras, Meha]ic, and

Burstadt (1971) [31] showed experimentally that as the

angular extent of the low pressure region was increased from

0 to 60 degrees, the loss in stability margin increased

drastically, but the loss was nearly constant between 90 and

180 degrees. The experiment also showed that somewhere

between 60 and 90 degrees, the prediction of loss of

stability margin from parallel compressor theory was more

reasonable. From these observations, a critical angle was

defined as the minimum distortion angle which would allow

prediction of stability margin loss by simple parallel com-

pressor theory.

Goethert and Reddy (1971) [32] arrived at a

critical angle of approximately 60 degrees based on a

theoretical treatment of a cascade of blades in a phase-

shifted oscillatory flow field such as would be produced by

a compressor rotor moving through a circumferentially dis-

torted flow field. In the unsteady cascade flow treatment,

it was shown that an airfoil (or cascade of airfoils) takes

a certain amount of time to change its lift in response to

rapid changes of angle of attack. It was also shown that

the change in the lift from the steady-state value during a

rapid oscillation of the approach flow direction diminishes

as the frequency of oscillation is increased. Thus, cir-

cumferential distortions with low pressure regions which

cover only small angular extents will be recognized by a

rotor blade passing through the region as a very rapid

15

A E D C-T R -77-80

change, and the change in lift will be small. As the

angular extent of the distortion is increased, however, the

rotor blade will have an increasingly longer time to adjust

to the higher angles of attack existing in the low pressure

region.

The portion of the compressor in the low pressure

region will therefore operate at angles of attack nearer the

stall value, thus reducing the stability margin of the

compressor.

The analysis of Goethert and Reddy was particu-

larly useful in that the difference in the response of a

cascade compared to that of an isolated airfoil was clearly

made. Also, the difference between oscillatory flow

approaching a stationary cascade and stationary flow

approaching an oscillating cascade was made. The influence

of unsteady airfoil aerodynamics on compressor behavior in a

distorted flow field was recognized and used by other

investigators also [33, 34]. Extensions have also been made

into unsteady airflow aerodynamics of stalled flow around

an airfoil by Carta (1973) [35] and Melick (1973) [36].

Uniform Time-Varying Entry Flow

The effect of uniform time-dependent disturbances

on the stability of compressors has also been investigated.

Gabriel, Wallner, and Lubick (1957) [37] studied the effects

of uniform fluctuations of compressor inlet total pressure

on compressor stability. The influence of rapid ramps in

16

AE DC-T R-77-80

compressor inlet total temperature was also studied.

Analysis of the influence was performed by solving a set of

lumped volume unsteady mass conservation equations along

with equations representing the pressure-ratio-mass flow

characteristic curves of several compressor stage groups.

The characteristic curves of the stage groups were obtained

from steady-state tests of the compressor and were assumed

to be valid dynamically in this analysis. This is the

so-called quasi-steady-state stage characteristic assumption.

The behavior of a fifteen-stage turbojet engine compressor

was analyzed by dividing the stages into four stage groups.

The resulting equations were solved using an analog computer.

It was assumed in the analysis that if any stage group

characteristic curve reached its peak value, that is, the

characteristic curve slope changes from negative to posi-

tive, that compressor instability would occur. The work

indicated that for uniform, sinusoidal oscillations of

compressor inlet total pressure, the amplitude required to

cause instability decreased as the oscillation frequency

increased. Also, it was found that upward ramps in com-

pressor entry total temperature would cause compressor

instability, and that the effects become increasingly

severe as the ramp rate is increased. The calculated

results were compared to experimental data and showed

reasonably good agreement with the experimentally determined

stability limit. This particular analysis was limited to

compressor rotor speeds near the design value.

17

AEDC-TR-77-80

Later investigators extended Gabriel, Wallner, and

Lubick's approach for calculating the effects on stability

of uniform, time-dependent compressor inlet disturbances.

Kimzey (1966) [38] improved the analysis by using each com-

pressor stage separately (instead of combining into stage

groups), and improved upon the form of the analog computer

solution. Further refinements, including the approximation

of unsteady momentum effects, unsteady energy effects,

improvement of the criteria for instability, and imple-

mentation of the solution on the digital computer, have also

been accomplished [24, 39, 40].

Time-Varying Distortion

Although a great deal of excellent experimental

work has recently been accomplished on time-varying

distortion [13 through 16], relatively little has been

accomplished to provide a method of directly calculating its

effects on stability. A quasi-steady approach, developed by

Plourde and Brimlow (1968) [7], is most commonly used at the

present time. This particular approach assumes that the

compressor reacts to time-variant distortion in the same

manner as it reacts to steady-state distortion if the time-

variant distortion pattern "dwells" at the compressor entry

for a time period approximately equal to the time required

for the compressor rotor to make one revolution. A dis-

tortion index, which correlates the loss in stability margin

to the spatial distortion pattern characteristics in the

18

AEDC-TR-77-80

same manner as is used with steady-state distortion, is then

calculated as a function of time. Time-dependent fluctua-

tions in the compressor entry flow field, which are at a

frequency higher than that commensurate with the pattern

"dwell time," are filtered out prior to making the dis-

tortion index computations. The loss in stability margin is

then correlated with the maximum distortion index value

observed over a given time period. Although Plourde and

Brimlow's treatment does not directly account for some

dynamic processes known to be present in the compressor with

time-varying distortion, reasonably good correlation is

achieved with the method, making it a most valuable engi-

neering tool for practical applications.

Melick (1973) [36] applied unsteady airfoil theory

to the analysis of the time-variant distortion problem.

Additional detail on time-varying distortion analysis is

provided in [41, 42].

V. OBJECTIVES AND APPROACH OF CURRENT WORK

As previously stated, the overall objective of

this study is to develop an improved analytical method with

general applicability for predicting compressor stability

margin loss caused by the aforementioned external dis-

turbances. The development of the method is intended to

provide insight into some of the physical processes which

lead to compressor aerodynamic instability.

19

AE DC-T R-77-80

The work described herein specifically extends the

present state of knowledge through the following develop-

ments.

Improved One-Dimensional, Time-Dependent Model

An improved model, suitable for steady-state

stability limit computation and for analysis of planar,

transient and dynamic disturbances is developed. The model

includes mass, momentum, and energy effects. The full non-

linear form of the equations, not used in earlier works, is

solved. Unsteady cascade influences, also not included in

earlier works, based on the Goethert-Reddy analysis [32],

are included in the analysis and their influences evaluated.

The Goethert-Reddy unsteady cascade analysis is compared to

experimental unsteady compressor blade aerodynamic data.

Detailed comparisons of model computed steady-state

stability limits are presented for three different com-

pressors providing an indication of the generality of the

analysis method. Four unsteady cases: compressor

instability caused by planar oscillatory inflow, dynamic

unstalled response of a compressor to oscillating inflow,

dynamic unstalled response to oscillating back pressure, and

response of a compressor to rapid inlet temperature ramps,

are computed. Where possible, the computed results are

compared to experimental results.

20

AEDC-TR-77-80

Extension to Multi-Dimensional Flow Case

The concepts of the one-dimensional, time-

dependent model are extended to three dimensions to allow

treatment of distorted flows into the compressor. The

compressor is divided both radially and circumferentially

into control volumes, and crossflows among the volumes

caused by distortion and their effects on stability are

computed. The influence of the radial work variation is

approximated and included for the radial distortion evalu-

ations. Unsteady cascade effects are included. Thus, a

model capable of computing the influences of combined radial

and circumferential distortion is produced. The model is

used to compute the loss in stability margin for a com-

pressor with steady-state, combined radial and circumferen-

tial distortion, and comparisons to experimental results are

made. A time-dependent distortion case is computed.

Parametric influence of pure radial and pure circumferential

pressure distortion are analyzed. Pure circumferential

temperature distortion is also analyzed.

Comparisons to Experimental Results

Emphasis was placed on comparisons of model

computations to experimental results to maximize under-

standing of the physical processes and to test the validity

of the computations. Experimental results on no single

compressor were sufficient for the comparisons desired;

thus, three different compressors were modeled. They are:

2!

AE DC-T R -77-80

a four-stage Allison XC-1 compressor [43], an eight-stage

NACA research compressor, the NACA-8 [44, 45, 46], and the

General Electric eight-stage J85-13 engine compressor [47].

Information on these compressors used in this study are

given in Appendix B. The NACA-8 and General Electric J85-13

compressors were modeled one-dimensionally and the models

included a portion of inlet ducting and combustor ducting

for more realistic time-dependent boundary conditions for

the compressor. The Allison XC-I compressor was used for

one-dimensional, steady-state stall analysis and for the

distortion studies. No ducting was included in the XC-I

models.

VI. ORGANIZATION

The development of the one-dimensional, time-

dependent model is presented in Chapter II. Chapter III

presents application of the one-dimensional analysis to

example problems. The extension of the model to three

dimensions is given in Chapter IV. Chapter V describes

application to distortion cases. The work is summarized

and recommendations for further work is given in Chapter VI.

22

AEDC-TR-77-80

CHAPTER II

DEVELOPMENT OF ONE-DIMENSIONAL,

TIME-DEPENDENT MODEL

The axial-flow compressor is made up of successive

rows of airfoils, a moving row, the rotor, followed by a

stationary row, the stator, as shown in Fig. 6. A rotor

followed by a stator is defined as a stage for this study.

A representative absolute velocity and static pressure

distribution through the compressor is also shown in Fig. 6.

The compressor increases the static pressure and density of

the fluid passing through it by the dynamic action of the

rotor which imparts kinetic energy to the fluid. Depending

on the specific design, the static pressure through the

rotor changes by varying amounts. The function of the

stator row is to diffuse and redirect the flow to convert a

portion of the imparted kinetic energy to internal energy.

The flow is further diffused as it passes from the com-

pressor through the compressor diffuser to the combustor.

After leaving the combustor it passes through a turbine

nozzle (if on an engine) or a throttling valve (if on a

compressor test rig) which is usually choked.

In the formulation of the one-dimensional model, a

control volume, enclosing the fluid in the compressor is

drawn, as shown by the dashed lines in Fig. 7. For the

one-dimensional analysis, all flow properties are "bulked"

AE DC-TR-77-80

in the radial and circumferential directions and variations

with the axial coordinate and with time only are considered.

The development of the one-dimensional model pro-

ceeds as follows. The governing equations are written. The

method of specifying the force and shaft work acting on the

fluid is developed. The finite difference forms of the

equations used for numerical solution on a digital computer

are then developed and the method of solution given.

I. GOVERNING EQUATIONS

Figure 7a illustrates the ducting-compressor-

burner-turbine vane system which was modeled. Figure 7b

shows the overall control volume representing the system.

The forces of the blading and walls of the system acting on

the fluid are represented by an effective axial time-

dependent force distribution, F(z,t). Similarly, shaft

work done on the fluid and heat added to the fluid are

represented by the time-dependent distributions, WS(z,t) and

Q(z,t). Mass removal or addition (interstage bleed, for

example) are represented by another time-dependent bleed

flow distribution function, WB(z,t). All bleed flows are

assumed to cross the control volume boundary normal to the

axial direction. Time-dependent boundary conditions are

provided by specification of the total pressure and total

temperature, P(t) and T(t), respectively, at the system

forward boundary and by specification of discharge static

pressure, Pexit(t), at the aft boundary. For choked turbine

24

AEDC-TR-77-80

vanes, a sufficiently low discharge static pressure was

specified to assure choking. In that case, the aft boundary

condition is unity Mach number. If unchoked, specification

of static pressure is the aft boundary condition.

The governing equations are derived by appli-

cation of time-dependent mass• momentum, and energy

conservation principles to the elemental volume of Fig. 7c.

Mass

Application of the mass conservation principle

(mass is neither created nor destroyed) to the elemental

control volume yields

BW B(DA dz) W+~ dz + WB dz + Bt = W Q

mass leaving time rate of mass entering control volume increase of mass control volume per unit time "stored" in the per unit time

control volume

(1)

Equation (1) may be reduced to

B(pA) = BW WB . ~t BZ

(2)

Momentum

The momentum principle, the summation of all

forces in a given direction, is equal to the time rate of

change of momentum in that direction, gives

25

AEDC-TR-77-80

F dz + P S A - P S A + ~(PS~zA) dz + P S A + ~ - ~ dz - A

axial forces acting " on control volume

wu a[wu) z] = +~--~----d - WU

morn en tu~ momentum leaving entering control volume control volume per unit time per unit time

+

time {ate of increase of momentum "stored" in the control volume

total time rate of change of momentum

(3)

Equation (3) may be reduced to give

~W 8(IMP) +F+PS ~A ~-~-- - ~z ~ '

(4)

where, the impulse, IMP, is defined as

IMP =WU+PS A . (5)

Energy

Energy conservation gives,

~H H+~ dz

enthalpy leaving control volume per unit time

+ ~t P ~- dz

time rate of increase of energy "stored" in control volume

= H

enthalpy entering control volume per unit time

+ WS dz + Q dz

shaft work heat added done on to fluid in fluid in control volume control volume

(6)

26

AEDC-TR-77-80

Equation (6) may be reduced to

where

a n d

(XA) _ ~t

~H ~-~+WS +Q , (7)

[ u2] X:p e+-f (8)

H = Cp WT . (9)

State and Additional Equations

Additional equations required include the perfect

gas equation of state,

PS = D R TS , (i0)

calorically perfect internal energy and enthalpy relation-

ships,

e=c v TS +constant , (11)

h = Cp TS + constant , (12)

stagnation (total) state temperature,

U 2 T=Cp TS +-~- , (13)

and stagnation (total) state pressure,

Y

, (14)

27

A E D C-T R -77 -80

where

7=cP CV °

Also required is the Mach number,

(15)

U M = -- (16) a r

where the acoustic velocity, a, is

a = ~7 R TS = ~y ~ , (17)

and the Mach number-static pressure form of the energy in

the control volume,

which results from combination of Eqs. (8), (10), (11),

(15), (16), and (17).

The Mach number-static pressure form of the

impulse, IMP, is also used,

IMP=PS A[I+TM 2] .

(18)

(19)

This equation results from combination of

W=pAU , (20)

with Eqs. (5) and (17).

28

A E D C-T R-77-80

II. STAGE FORCES AND SHAFT WORK

Steady-State Sta@e Characteristics

Consider the element of fluid contained between

two blades in a rotor as shown in Fig. 8a. Entering the

rotor, the tangential velocity is V 8 , and at the exit it is 1

V82. Applying the conservation of angular momentum

principle to the fluid element in steady flow gives the

t=W[r2 V8 -r I ] • 2 VSl

torque on the element,

(21)

Multiplying Eq. (21) by rotor speed, ~, yields the time rate

of doing work on the fluid, which is equal to the total

enthalpy change across the rotor. Thus,

o~ [, ,i [r r ] W - = ~W V@2 V@2 (22)

Replacing the difference between radius-velocity products

using an effective radius yields,

opwE,2 ,l~ o~[vo2 vol]

= -r 1 r2 V82 V81

V82 -V@I

where

(23)

(24)

29

A EDC-TR-77-80

Writing wheel speed as

and

Uwh = ~r ( 2 5 )

gives

- = AV 8 (26) V82 V81

T 2 -T 1 = Uwh AV 8 •

T 1 Cp T 1 (27)

Applying the linear momentum equation for steady

flow to the element of Fig. 8a gives

F 8 =W AVe =Pl A1 U 6V 8 , (28)

where F 8 is the force being applied to the fluid by the

blading in the compressor.

The force, F8, can also be expressed in terms of

blade lift and drag, as indicated in Fig. 8b,

F 8 =L cos(e+k) + D sin(e+A) , (29}

where u is the blade angle of attack, determined by the

velocity triangle of Fig. 8b, and X is the blade chord-to-

axial direction angle, or, the cascade stagger angle.

30

AEDC-TR-77-80

ficients,

and

The lift and drag may be expressed as coef-

L C£ =i V 2 ' (30)

Pl rel Aref

D Cd = 1 Pl V2 tel Aref

(31)

Combining Eqs. (27) through (31) produces

• 1" 2 - T 1

T 1

[ 1 A__~f] Uwh~rel [ ] (32)

Returning once more to Fig. 8b, define the stage flow

coefficient as

U --- . (33}

V e

From Fig. 8b, the flow coefficient, %, is related

to the blade angle of attack, u, by the stagger angle, A,

which is a constant for any given blade row,

¢ = cot(A+a) . (34)

Thus, ~ serves the same purpose in the compressor as ~ does

for aircraft wing aerodynamics. It decreases with

increasing u; therefore, small ¢ means large a, and vice

versa. Also from Fig. 8b,

31

AEDC-TR-77-80

U - - = c o s ( X + ~ ) Vrel

(35)

Combining Eqs. (32) through (35) yields

% ~-i i~ u 2 v 2

T 1

(36)

The left-hand side of Eq. (36) is defined as the stage

loading parameter, #T (following Horlock [20]), or simply as

the stage temperature coefficient (more common in American

works [40]). Thus,

~T = cp(TR-I) ,

U 2 wh T 1

(37)

where TR is the stage total temperature ratio.

If the swirl component of tangential velocity is

smal i,

and

U = - (38)

Uwh

= 2 A 1 + 1 C£ + ( 3 9 )

32

AE 0C-TR.77-80

At given Mach and Reynolds numbers, C£ and C d are functions

of u; thus, #, only. Therefore,

~T I = T(#) . (40) Re,M

In principle, theoretical or experimental lift and

drag coefficients could be used, along with stage geometry,

(A, AI, Are f) to compute the stage temperature coefficients

following Eqs. (36) or (39). Corrections for blade row

interference causing losses and secondary flows would be

required. Another method is to measure the stage total

temperature, flow rate, total pressure, and flow angularity

at the stage entry and exit and directly compute ~T and

from Eqs. (37) and (38). The stage characteristics used in

this study were obtained in that manner and are summarized

in Appendix B.

The temperature ratio of Eq. (37) for an ideal

compressor stage is the isentropic temperature ratio defined

from stage entry and exit stagnation pressure. Using this

ratio, another stage parameter, the stage pressure coef-

ficient, ~P, is produced,

~-__!i

2 Uwh

T 1

(41)

33

AE DC-TR-77-80

The pressure coefficient is also a function of the flow

coefficient, ~, in a manner similar to the temperature

coefficient. The stage adiabatic efficiency is

had = Ahisentropic = PR 7 -1

Ahactual TR-I (42)

Thus,

~P ~ad = V "

(43)

A set of curves,

CT = %T (¢)

I i

~ad = had (~) I J

, (44)

is referred to as stage characteristics, and along with flow

direction,

[%]. (45)

or an equivalent set of flow direction information, fully

defines a stage's performance from a one-dimensional stand-

point. Note that the set of Eq. (44) is redundant; any two

of the three variables are sufficient to compute the third.

Various forms of the stage characteristics are

used. An alternate form used in NACA compressor research,

34

AEDC-TR-77-80

and used for the stage characteristics of the NACA-8 com-

pressor treated in this study, is the equivalent temperature

and pressure ratio form. They are defined as

LCp T J design

EPR = ~P + 1 , (47) LCp T J design

and had is defined as in Eq. (43).

Because

CpTJ design

is a constant for a given compressor, the set remains a

function of ~, as in Eq. (44), and contains the same

information.

Figure 9 is an example set of stage character-

istics for the first stage of the NACA-8 compressor [46].

To the right of the equivalent pressure ratio peak, the

stage is unstalled (i.e., high ~ and low ~). To the left,

the stage is stalled. Rapid decrease of stage efficiency at

low ~ (high u) where the blades are stalled, is evident.

The data of Fig. 9 is seen to generalize well with the flow

coefficient, ~, computed as in Eq. (38) with the swirl

component of velocity ignored. This implies that for this

set of data, the swirl velocity component was either small

35

AE DC-TR-7740

relative to wheel speed, or, changed little as a fraction of

wheel speed over the range of compressor operation con-

sidered. Steady-state stage characteristics for the three

compressors treated in this study were obtained from [43,

46, 47], and are shown in Appendix B.

The foregoing discussion pertains primarily to the

compressor rotor which imparts the shaft work to the fluid.

No stagnation temperature change occurs across the stators

(with adiabatic boundaries) and the stagnation pressure

losses of the stator may be combined with the change across

the rotor to give a net change. Therefore, the character-

istics of Eq. (44) may be attributed to the overall stage.

Unsteady Cascade Effects

When the angle of attack of the flow into the

compressor stage varies slowly, the lift coefficient follows

the steady-state C£-u curve very closely, as illustrated in

Fig. I0. Similarly, the steady-state ~-~ characteristics

are followed. When ~ rapidly changes, however, the lift

cannot instantaneously follow because a certain amount of

time is required for the flow around the blade to adjust to

the new condition. Lag loops or hysteresis in the CE-~ and

~-~ curves appear (Fig. 10).

Goethert and Reddy analyzed the unsteady cascade

effects [32]. They analyzed unsteady incompressible

potential flow through a stationary cascade of thin, flat

plate foils with small oscillations in the axial velocity

36

AEDC-TR-77-80

producing angle-of-attack oscillations. The time-averaged

angle of attack was zero. Their results are summarized in

Fig. ii. The cascade stagger angle is A; the blade chord

length, C; and spacing is a. Rotor speed, ~, was constant

and Vsw was zero. The axial velocity, U, was oscillated at

frequency f, causing an inverse oscillation of ~. This

produced an oscillation in the lift coefficient. The

results were presented in terms of maximum amplitude ratio

at a given frequency, divided by the lift coefficient

amplitude at f = 0 (quasi-steady-state). The resulting lift

ratio is presented as a function of reduced frequency, k,

for four values of cascade spacing, a/C. The case of

a/C = =, a single, isolated airfoil, is shown for reference.

Also treated for reference was the case of oscillating foils

in a steady stream. The Goethert-Reddy work showed signifi-

cant differences to exist between the oscillating foils,

stationary stream, and the converse case. Their work also

showed the unsteady effects for a cascade of finite spacing

to be much less severe than that of an isolated airfoil.

The unsteady influence was shown to increase with increasing

spacing (a/C) and stagger angle, A.

The lift coefficient ratio variation with reduced

frequency is intuitively correct at k = 0 and as k ~ ~.

Following Goethert and Reddy [32] the reduced frequency, k,

may be viewed as an indication of the ratio of flow passage

time to the disturbance stay time of the forced oscillation.

37

A E DC-TR -77-80

The passage time is defined as the time required for an air

particle to pass through the blade row channel. Or,

_ C ( 4 8 ) Atpassage Vre I "

The disturbance stay time is defined as the time that any

one blade is exposed to the positive disturbance velocity of

one velocity oscillation, that is,

Therefore,

1 Atdisturbance = 2-~ " (49)

dt~assa~e = 2f___.~C=

Atdisturbance Vrel (50)

Alternately, k may be viewed as being proportional

to the ratio of the blade chord length to the disturbance

wave length,

chord length _ C _ Cf _ k wave length Vrel Vrel ~-~ . (51)

If k is near zero (i.e., the disturbance time period is

large compared to the passage time), the blade lift will

easily follow the disturbance. Similarly, the chord length

is small relative to the disturbance wave length; thus, the

lift can easily respond. If k * ®, the disturbance time is

small compared to the passage time and many disturbance wave

38

A E D C-T R-77-80

lengths reside within one chord length at any instant in

time. The disturbance is then "averaged-out" by the blade

yielding an instantaneous lift value of zero due to the

disturbance.

Bruce and Henderson [48] conducted an experi-

mental investigation of the unsteady force coefficients

produced by a compressor rotor in unsteady flow. A single

stage, low-speed compressor was fitted with a rotor blade

having a segment of the blade on a small strain-gaged force

balance. This installation allowed a direct measurement of

the unsteady forces acting on that blade segment as the

rotor passed through a distorted flow field produced by a

distortion screen. A portion of their measurements was

taken with stagger angle and cascade spacing near the values

used by Goethert and Reddy. Figure 12 shows a comparison of

the Goethert-Reddy analysis with the Bruce-Henderson experi-

mental results. The trends with reduced frequency are in

good agreement. The experimental results generally were

from I0 to 20% higher than the analytical results. Con-

sidering the difficulties and assumptions involved in both

the analytical and experimental evaluations, this level of

agreement is considered reasonable at this time. Agreement

among other theoretical treatments and experimental results

surveyed was usually of this order or worse. (See compari-

sons made in [48] for example.)

The analysis of Goethert and Reddy was performed

for an unloaded cascade, that is, the mean angle of attack

39

AE D C-T R-77-80

was zero. Figure 13 shows the variation of lift coefficient

ratio with mean angle of attack from the unsteady cascade

work of Ostdiek [49], and Bruce and Henderson's single-stage

compressor [48]. The dynamic lift coefficient ratio magni-

tudes were normalized by the value of a mean angle of attack

of eight degrees. The measurements were made over a range

of reduced frequency of 0.002 to 2.9, and from 0 to 12

degrees mean angle of attack. The results are all within

±10% and no generalized trend could be identified.

Unstead~ Stave Characteristic Correction

The Goethert-Reddy analysis may be used to con-

struct a correction to the steady-state stage character-

istics. To facilitate application of the correction, the

dynamic lift coefficient ratio was represented by a simple,

first-order differential'equation, as indicated in Fig. 14.

The equation,

dC£ C£ C£s s + - (52) T T '

represents the time-dependent lift coefficient, C£, in

response to the quasi-steady-state lift coefficient, C£ss,

which would be produced if the lift exactly followed the

steady-state C£-~ relationship.

The blade time constant, T, depends on the chord

length and reduced frequency,

40

A E DC-T R-77-80

EC "[ - • (53)

Vrel

The constant, E, depends on cascade geometry (i.e., a/C and

A). For a/C = 1.0, and ~ = 50 degrees, £ = 0.3106. The

value of E is found by fitting the approximating differ-

ential equation, Eq. (52), to the response indicated by the

Goethert-Reddy analysis.

For sinusoidal variation of angle of attack, the

amplitude response of Eq. (52) is

IC~Imax I , = 1

Ic~ls~ + (~ ~) . (54)

The approximation of Eq. (52) is compared to the Goethert-

Reddy solution in Fig. 14, and fidelity .is excellent up to

k=8.

The cascade airfoil lift coefficient is related to

the stage characteristics through Eq. (39). Consider

oscillations of the flow into a stage around a given stage

characteristic operating point (~o' ~o )" Let ~ represent

either ~T or ~P. Expanding Eq. (39) in a Taylor series,

neglecting the drag coefficient term and retaining only the

first-order terms of the series gives

11'2 ~" A 1

(55)

4]

AEOC-TR-77-80

Near the point, (~o' ~o )'

- ~O

- ~o 0

(56)

Then, from Eqs. (55) and (56),

'to -1

Are f 1/2 - ] . (571

Substituting Eq. (57) into Eq. (52) produces,

d~ ~ ~ss EE + E = - E - • (58)

Thus, the stage characteristics may be corrected for

unsteady cascade airfoil influences using the same first-

order differential equation algorithm as that constructed

for the unsteady lift coefficient.

The linearization of the unsteady correction

implicit in Eqs. (55) and (56) is justified by: i) the

Goethert-Reddy analysis is itself a linearized analysis;

2) the correction magnitude is small in the range of reduced

frequency normally encountered, as will be demonstrated

later. The correction is actually proper only up to the

stall point (i.e., up to the point where unsteady viscous

effects become predominant).

42

AEOC-TR -77-80

Force and Shaft Work Distributions

The momentum equation, Eq. (4), requires an axial

force distribution representing the force of the compressor

blading and casings on the fluid. The function is con-

structed from the experimentally determined steady-state

stage characteristics corrected for unsteady effects. (Or,

alternately, the force may be computed in a quasi-steady-

state manner from the steady-state stage characteristics and

then directly corrected for unsteady effects using the

dynamic lift coefficient ratio of Fig. 14.) The shaft work

distribution required by the energy equation, Eq. (7), is

similarly constructed.

The exact method of extracting the force and shaft

work distributions depends partly on the finite difference

method used to solve the governing equations, and is

described in the next section.

Ill. METHOD OF SOLUTION

The basic governing equations are:

(~A) ~W ~----~ = ~z WB , (2)

and

~W _ 3 (IMP) + F + PS 5A 9t ~z B--{ , (4)

(XA) _ ~H Bt ~z + WS + Q . (7)

43

AEDC-T FI-77.80

The dependent variables to be integrated for are pA, W, and

XA. The area, A, was preserved in the derivative to allow

for cases where exit area is rapidly changed.

Finite Difference Approximation, One-Sided

Difference Form

The spatial derivatives were approximated in two

different finite difference forms. When only the compressor

was considered Ino ducting), a simple, one-sided difference

approximation was used. The control volume was divided into

unequal lengths. Each length included a stage of the com-

pressor as shown in Fig. 15. Figure 15 also shows the

computation plane designations. The spatial derivatives at

location i were approximated by

aw. w. - w. i ~ 1 im

-~z zi _ Zi m , 1591

a (IMP) i

az

~ IMPip - IMP i

Zip - z i , (60)

aA. 1

-~ = constant , 161)

(the product of static pressure with the area-derivative was

combined with the force term, F),

and

aH i H i - Him m

-~z z. - Z. " (62) 1 im

44

A E DC-TR-77-80

Substitution into Eqs. (2), (4), and (7) produces

the approximate equations,

~(PA)ip = Wi - Wip

8t Zip - z i ' (63)

and

~w. IMP. - IMP.

_ i ~ P + F~ (64) -~t z. - z. J. ap i

(XA) . H. - H. ip = i ip + WS. + Qi (65)

~t z. - z. 1 " ~p

Writing Eqs. (63), (64), and (65) from i = 1 to

i = in-l; that is, from the forward to the aft control

volume boundaries, results in 3x(in-l) first-order differ-

ential equations to be solved simultaneously.

Force and Shaft Work Computation

The division of the control volume into stage

lengths was made because the forces and shaft work applied

between stage entry and exit are obtainable from the stage

characteristics. To determine the force and shaft work

applied to each element, the flow coefficient into the

element is calculated,

U • = 1

~i ~ . (66)

1

45

AEDC-TR-77-80

The stage characteristic, represented as a polynomial curve

T (or n i) are found. fit, is then entered and ~ and ~i

Corrections for unsteady effects are made. The stage stag-

nation pressure ratio and temperature ratios which go into

the force and shaft work computations are calculated from

a n d

wh P PR i = = 1 + c~ ~i (67)

wh T TR i = = 1 +c~ ~i "

Shaft work is computed as

WSi = WiTi[ TR i - i] .

Similarly, the axial force is given by

(68)

(69)

= IMP [IMPR. - 1] (70) Fi i I "

IMPR i is the impulse function ratio across the stage,

determined from the stagnation temperature and pressure

ratios, area ratio, and flow directions,

Y

46

AE D C-T R-77-80

The variable, ME, is the equivalent Mach number downstream

of the stage which would be present if steady flow existed.

(Recall that the stage characteristics from which the forces

and shaft work are determined, are based on experimental

steady-state measurements). The value of ME is computed

from a polynomial expression approximating the inverse of

Eq. (72), for constant R and 7,

r- +ll I y---l-I

(72)

MFFE is the downstream equivalent mass flow function,

w VT i TR i MFFE = i (73)

Pi PRi Aip cos Bip '

defined in the same spirit as was ME.

The flow direction angles, ~i and Sip, taken with

respect to the axial direction, are input as functions of

compressor corrected rotor speed, as constants, or, are

assumed to be zero if no flow direction information is

available.

Solution Procedure

Figure 16 outlines the overall digital computer

solution procedure followed. The problem is solved as an

initial condition problem. Following the flow path of

Fig. 16, at time equal zero, initial values of the dependent

47

AEDC-TR-77-80

variables are specified. Entry plane, interior plane and

exit plane conditions along with forces and shaft work are

computed for time equal zero and are used to compute the

time derivatives of the dependent variables. The time

derivatives are then integrated using a fourth-order Runge-

Kutta numerical integration method [50] for the dependent

variabie values at the next step in time. The new values

are then used to compute conditions for that time step and

the solution advances in time until the desired solution is

obtained.

Some additional details of the computation pro-

cedure are as follows.

Entry plane conditions. At the forward boundary,

i = 1, flow, Wl, will be available from integration of

Eq. (64). Density, PI' and the energy variable, Xl, are

not available and must be supplied by the boundary condition.

Stagnation pressure, PI' and stagnation temperature, TI, are

supplied as boundary conditions. The mass flow function,

is computed.

function by

wi 7 MFFi = PIA1 ' (74)

Mach number is related to the mass flow

y--rlj (75)

48

AE D C-T R-77-80

The inverse form of Eq. (75) is approximated by a polynomial

curve-fit of M 1 as a function of MFF 1 for constant R and 7.

Static temperature, static pressure, density, and the energy

variable may be computed knowing Mach number:

-1

[ Tsz ] PSz = PI L ~?z J

(76)

, (77)

and

PS 1

Pl - R TS 1 '

Psl[ 21 X 1 = ~-~ 1 + ~ M 1 .

(78)

(79)

Interior plane values. On the interior planes,

values of the three dependent variables are known from

integration. To provide the other variables necessary for

the solution, the following calculations are made at each

W, 1

Ui = pi-~--~i '

plane:

(80)

_ 1[xi TSi Cv Pi -~ ' (81)

PSi = Pi R TS i , (82)

49

AEDC-TR-77-80

u 2

T i = Cp TS i + ~ , (83)

Pi = PSi T~ i

7

, (84)

a i = ~7 R TS i , (85)

and

U, M. - 1 1 a i '

IMPi = PSi Ai[l + 7 M2]z

(86)

, (8";)

H z. = Cp W.~ T i . (88)

Exit plane values. Density, p, and the energy

variable, X, are available from integration. Flow, W, is

not available from integration and must be found using the

aft boundary condition. At the exit plane of the aft

control volume, static pressure is specified. Exit

stagnation-to-static pressure ratio is tested against

critical pressure ratio to determine if the exit is choked

or not, as follows:

7

[ ]ori. (89)

50

AEDC-TR-77-80

If

[P] Pxn • ~

Pexit -- crit

the exit is choked and

Min = 1.0 .

Then,

If

(7-1) Xin PS. =

~n i + 7(7-1) " 2

(90)

Pexit crit

the exit is unchoked, and

PSin = Pexit "

Therefore,

q[ 2 Min =

Xin 1 ]

PSin 7-i (91)

Mach number and static pressure are now known for either

condition. Therefore, the remaining calculations are conuoon

for either case.

5]

AED C-TR-77-80

and

PS. in TS. -

in R -1°-n ' (92)

Uin = Min ~ 7 R TSin , (93)

Win = Pin Uin Aexit " (94)

Impulse and enthalpy values at the exit and entry

planes are computed using Eqs. (87) and (88).

Finite Difference Approximation, Central

Difference Form

When ducting ahead of and behind the compressor

was included in the system modeled, a noncentered, two-sided

difference form of finite difference approximation was used

to allow more accurate representation of the duct flow

processes and to avoid some numerical stability difficulties

experienced in attempts to apply the one-sided difference

form to duct-flow cases.

The central difference algorithm used is derived

as follows. Let y(z) be any variable with continuous third

derivative with respect to z, for which an approximation to

o~z desired. Referring to Fig. 15, let is

and

~Zip = Zip - z i ,

dZim = Zim - z i •

(95)

(96)

52

AEDC-T R-77-80

Expanding around the i th location,

a n d

i

.~j. . ~I .:°.0[.:.] Yim = Yi 8z AZim + 2 8z 2 li

(97)

(98)

Multiplying Eq. (97) by AZ~m, Eq. (98) by AZ~p, differencing the results and solving for ~-~zli gives

where

~z] -a +a.,.+ -0[~z'] i im Yim aip Yip

Az=max [AZip , AZim] ,

and the weighting terms are

• (99)

(100)

and

AZ. a. = ip

am Az. am AZip - AZim ]

a i = -

dZim + ~Zip AZip dZim

AZ. am 1 aip'~zip [~,io_~zip]

I (i01)

53

AEDC-TR-77-80

The sum of the coefficients,

aim + a i + aip 0 (102)

Multiplying Eq. (102) by Yi and subtracting from Eq. (99)

gives

(lO3)

Replacing the spatial derivatives in the governing

equations, Eqs. (2), (4), and (7), produces the alternate

finite difference form,

~[~,]~ "i~[ ~ -,~ip ] - ~ , (104)

~W.

@= aim [ IMP i - IMPim] + alp [ IMP i - IMPip ]

and

~A +F+PS. 0

i (105)

~[~,]~ ~ ° o~ [ ~ - ~ ] * ~ [ " ~ - . ~ ] . ~ . o . (106)

As was done with the one-sided case, the pressure-area

partial derivative in Eq. (105) was combined with the axial

force term.

54

A E D C-T R-77 -80

The solution procedure followed (Fig. 16) was the

same as used in the one-sided case except Eqs. (104), (105),

and (106) were used instead of Eqs. (63), (64), and (65).

The bleed flow, forces, shaft work and heat added

must also be made compatible with the two-sided difference

form. Thus,

and

WB = aim WBim -aip WB i , (107)

F= -aim Fim+aip F i , (108)

WS= -aim WSim+aip WS i , (109)

Q = - aim Aim + aip Qi " (110)

Also, on the forward and aft boundary planes, an appropriate

one-sided form of the differential equation was used. The

force term for ducting was taken as a simple friction loss

expressed as

F i = -Cd. 7 PS i A i M~I " (iii) 1

The value of Cd. was adjusted to give a reasonable stag- 1

nation pressure loss in the ducting and to permit

numerically stable operation of the model.

55

AEDC-TR-77-80

Interpretation and Limitations of

Differencin~ Methods

Differencin~ al@orithm truncation error. The one-

sided algorithm is a first-order approximation (i.e., it

ignores terms of order (~z) and above). The central differ-

ence algorithm is a second-order approximation (terms of

order (dz) 2 and above are ignored).

Let y again be any variable in the problem

requiring spatial differentiation. The one-sided algorithm

is found from expansion around a point,

i x 2 ~z2 i (112)

Solving for the first partial derivative,

i

(113)

The truncation error in the governing equations is the error

resulting from the substitution of the algorithm for the

exact form of the spatial partial derivative. Therefore, an

upper estimate for the truncation error for the one-sided

case is

1 ~2y, ~ -- ~- ~-~-~-! (~z) .

oZ 'i (114)

56

A E D C-T R -77 -80

The second partial derivative is finite, so, as Az -7 0, the

error vanishes and, referring to Eq. (112), the approxi-

mation is consistent of order one.

A similar analysis of truncation error for the

central difference algorithm produces an error estimate of

E ~ ~ (nz) .

~z i

(ll5)

The central difference algorithm is therefore consistent of

order two.

For the problems solved using the nonlinear

system of equations produced by the approximation, numerical

stability was investigated by running various time step

sizes and observing the behavior of the solution at large

times.

The system of equations constituting the model is

for time-dependent, one-dimensional subsonic flow and is

hyperbolic in nature, with specified boundary conditions and

initial conditions. Thus, each integration step to compute

the dependent variables at the new time, (t + At), must be

made within the region of influence of the known values at

time, t. The Courant-Friedrichs-Lewy maximum time step

criteria [51],

dZmin (116) Atmax - a + Uma X '

57

A E DC-TFI-?7-80

must be observed. For AZmi n = 0.25 ft, a = 1553 ft/sec

(value at TS = 1000aR), Uma x = a,

At = 8x10 5 sec . max

Values used were typically one-half of dtma x. Various

values, depending on the nature of the specific analysis,

were used.

With consistency and numerical stability, con-

vergence is implied [51]. The best proof of convergence of

the numerical solution is provided by the comparisons to

experimental results given in Chapter III.

Disturbance fre~uenc~ limitations. The differ-

encing length is Az; therefore, no disturbances having

disturbance wave lengths of order Az or less can be

explicitly treated by the model. As a working limit,

minimum wave lengths were restricted to i0 AZma x. There-

fore, the highest frequency phenomena with downstream

propagation of pressure disturbances which could be

explicitly treated by the model would be

(U + a)mi n = , (117)

fmax 10 AZma x

where U is a representative minimum axial velocity, and a is

the acoustic velocity. For U = 300 ft/sec, a = 1116 ft/sec,

and AZma x = 0.25 ft,

58

A E D C-T R -77-80

f = 566 Hz . max

Similar limits can be computed for temperature rather than

pressure oriented disturbances.

Steady-state solution. Inspection of the approxi-

mating equations, Eqs. (63), (64), and (65), or Eqs. (104),

(105), and (106), shows that when the time derivatives are

zero, the remaining equations are the exact mass, momentum

and energy equations for steady, one-dimensional flow

through a control volume of any specified size. The

governing equations in the approximate form could also have

been derived from an integral analysis of time-unsteady flow

through a finite control volume.

59

AEDC-TR-77-80

CHAPTER III

EXAMPLE PROBLEMS WITH COMPARISON TO EXPERIMENTS,

ONE-DIMENSIONAL MODEL

Three steady-state examples and three dynamic

examples are given. Three different compressors, the

Allison XC-I, the NACA-8, and the General Electric J85-13,

were driven to their stability limit for steady flow to

demonstrate the physical processes and to test the models'

ability to compute the steady-state stability limit. The

NACA-8 compressor was stalled dynamically by imposed

oscillating inflow. Dynamic response of the J85-13 for

first oscillating inflow and then for oscillating back

pressure was computed. Finally, the response of the NACA-8

compressor to a rapid inlet temperature ramp representing

rocket exhaust gas ingestion is presented.

Three different compressors were modeled because

experimental data necessary for comparison to the analytical

model did not exist in a sufficiently complete set for any

single compressor. A side benefit of modeling three com-

pressors is the demonstration of the generality of the

analysis method.

50

A E D C-T R-77-80

I. STEADY-STATE STABILITY LIMITS

The Allison XC-I compressor is a four-stage

compressor with design airflow and stall pressure ratio of

38.7 lbm/sec and 4.9:1. Details of the compressor, the

overall compressor map, and the stage characteristics used

in the model are presented in Appendix B. The information

on this compressor was taken from [43]. The computational

stations for the model are shown in Fig. 17a. No ducting

was included in this model, thus, the one-sided difference

formulation was used.

The compressor was loaded to the stability limit

by increasing the exit pressure until flow breakdown was

indicated. Inlet stagnation pressure and temperature were

constant. Corrected rotor speed was 90% of design.

Figure 18 illustrates the loading and the resulting stall

process. Exit static pressure was ramped at a rapid rate to

minimize computer time necessary for the computations. The

rate was sufficiently slow, however, that transient effects

were very small. Airflow at any stage never differed from

compressor inlet airflow by more than 0.5% during loading.

As the load on the compressor increased, compressor pressure

ratio increased and airflow decreased until flow breakdown

was indicated. Figure 19 shows some key compressor

variables on an expanded time scale covering the flow break-

down as computed by the model. Stage 3 flow coefficient

reaches its stall value at t = 55.7 msec. The flow, W3,

61

A E D C-T F1-77 -80

into stage 3 is seen to begin to decrease at the same time.

Shortly afterward the flows at each location begin to

diverge one fr~n the other and to oscillate. The static

pressure behind stage 3 rapidly falls because the stage's

pumping capacity is quite limited when stalled. The

pressure forward of stage 3 rapidly increases in response to

the sudden blockage to the flow which the stalled stage 3

produced. This static pressure signature is observed

experimentally, and is a common means employed to determine

which stage in a compressor stalled first, causing the

instability to occur. (See Fig. X-27 in [14] for example.)

Figure 20 shows the 90% corrected rotor speed

characteristic computed from the model compared to experi-

mental measurements. The flow breakdown point indicated by

the model agreed well with the experimental stability limit.

The model solution thus indicates the following

sequence leading to compressor instability. The compressor

is loaded and flow reduces until a stage somewhere in the

compressor reaches its stall point (its maximum pressure

coefficient point). The ability of that stage to continue

to pump flow (i.e., to help support the pressure gradient

over the length of the compressor imposed by the back

pressure load) begins to diminish. As the flow reduction

because of the increasing load continues, the ability of the

stage to pump diminishes even further because of the

increased load. Eventually, a point is reached where the

flow into the stage becomes fully stalled, thus, little

62

A E DC-T R -77-80

force on the fluid is provided to pump the flow, and the

flow path through the compressor becomes blocked because of

the highly separated condition of the stalled flow in the

channels. In terms of the solution of the model equations,

as the loading is increased, a point is reached where no

stable solution to the governing equations exists which will

satisfy the imposed load (boundary conditions); thus, the

solution, like the flow through the compressor, becomes

unstable. However, the behavior of the model quickly

becomes nonrepresentative of the compressor shortly after

flow breakdown because the stage characteristics are not

valid for deeply stalled, dynamic situations.

NACA-8 Compressor

The NACA-8 compressor is an eight-stage com-

pressor with design airflow and stall pressure ratio of

65 lbm/sec a~d 10:1. Compressor details and stage charac-

teristics are given in Appendix B and are taken from [44,

45, 46]. The model included inlet ducting and combustor

ducting, thus, the central difference form was used.

Computation planes are shown in ~ig. 17b. Figure 21

illustrates the loading of the NACA-8 model to the stability

limit by reducing the turbine nozzle area. The nozzle was

choked. Corrected rotor speed was 90% of design, inlet

stagnation pressure and temperature were constant. Pressure

ratio increased and corrected airflow decreased until

instability was encountered. The process was repeated at

63

AE DC-TR-77-80

70 and 80% corrected speed. The results are presented in

Fig. 22 and are compared to experimental data. Reasonable

agreement was obtained.

Extensive internal steady-state experimental

measurements were made on the NACA-8 compressor allowing

detailed model-to-experiment comparisons. Figure 23 shows

the variation of stage flow coefficients near the stability

limit as a function of corrected rotor speed. Both experi-

ment and model computed values are shown. Good agreement

was obtained.

It is also evident from Fig. 23 that this com-

pressor operates with forward stages below the stall value

of stage flow coefficient at corrected rotor speeds below

approximately 78%. Both experimental and computed flow

coefficient values for stage 1 are below stall at 70% speed.

This behavior is in agreement with previous theories and

experiment observations (see Chapter I and [21, 24]). An

explanation is as follows. At low corrected rotor speeds,

the forward stages are unable to deliver the design pressure

ratio. The flow density provided to the aft stages is

therefore much lower than that for which the compressor

axial area distribution was designed. Thus, to maintain

steady mass flow continuity, the flow coefficients (axial

velocity) in the aft stages are high. Indeed, at suffi-

ciently low corrected rotor speeds, the aft stages are even

choked. This limits the flow through the compressor to a

low value. It also means that the aft stage characteristics

64

A E DC-T R-77-80

will be steep relative to the forward stages (see Appendix

B). Therefore, at low corrected rotor speeds, as the load

on the compressor (the demanded pressure ratio) is increased,

flow reduces, and the following sequence occurs. The

ability of the aft stages to increase their pressure ratio

in response to the flow reduction is initially greater than

the decrease in pressure ratio occurring in the forward

stages in response to the same reduction in flow rate. Thus,

the increased load may be satisfied. As the loading con-

tinues, however, the forward stages become more and more

stalled, and a point is eventually reached where the aft

stages can no longer compensate. At that point, the load

can no longer be satisfied, and overall compressor insta-

bility occurs.

This example makes clear the need to distinguish

between stage stalling and overall compressor stalling or

instability. Overall compressor instability is always

preceded by stage stall, but is not always simultaneous with

stage stall.

General Electric J85-13 Compressor

The General Electric J85-13 compressor is an

eight-stage compressor with design airflow and stall

pressure ratio of 43 ibm/sec and 7.7:1. The compressor

employs a set of variable inlet guide vanes ahead of the

first-stage and compressor bleed from the second-, third-,

and fourth-stage stator cases. The guide vanes and bleed

65

AE DC-TR-77-80

valve are scheduled by the engine control as functions of

corrected rotor speed. (Additional bias for inlet tempera-

ture is also sometimes provided.) The schedules used in the

model of this study are given in Appendix B. Stage charac-

teristics and additional compressor details are also given

in Appendix B and are taken from [47].

The steady-state stability limit and corrected

speed characteristics were computed for 80 and 87% corrected

rotor speed and are compared to experimental results

(Fig. 24). Agreement was good. The compressor model was

loaded by reducing turbine nozzle area. The nozzle was

choked. The model included ducting, thus the central

difference formulation was used.

II. DYNAMIC STABILITY LIMITS

Oscillatin~ Inflow Loading, NACA-8 Compressor

Figure 25a shows the oscillation of the inlet

total pressure imposed on the NACA-8 compressor model.

Inlet total temperature and turbine nozzle area were

constant. Corrected rotor speed was 70%. The pressure was

oscillated at 20 Hz and the amplitude was slowly increased

until compressor instability occurred. Overall compressor

pressure ratio responded as shown in Fig. 25b. For con-

venience, the model was started at 90% speed (initial

conditions were previously calculated for 90% speed),

rapidly decelerated to 70% speed, stabilized, and then the

oscillation was imposed at t = 0.15 second.

66

AED C-TR-77-80

The pressure ratio oscillations grew with the

imposed pressure oscillations until instability was

encountered. The response of the airflow into the com-

pressor is shown in Fig. 25c. Initially, the flow variation

is near sinusoidal, but, as the stages are driven into and

out of their stalled regions, indications of flow breakdown

begin to appear. Finally, the flow completely diverges to

zero. The phase shifts and amplitude variations of the flow

properties through the compressor due to the forced oscil-

lations cause the stage flow coefficients to oscillate.

Figures 25d through 25g show the flow coefficient dynamic

variations. Also indicated are the individual stage stall

values and the value of each flow coefficient at which the

compressor steady-state stability limit was reached at 70%

rotor speed. Excursions well below both values occur

dynamically before total flow breakdown is experienced.

This behavior is in qualitative agreement with experimental

observations reported in [5, 37]. Examinations of the flow

coefficient variations indicate that stage 7 was the stage

which eventually became totally unstable and caused com-

pressor instability (Fig. 25g).

Interpretation of the model results suggests the

following physical process for instability with oscillating

inflow. The stage flow coefficients are alternately driven

into and out of their stalled regions. The stages must be

held in the adversely loaded condition for a sufficiently

long period of time for the flow breakdown to occur. Thus,

67

AE DC-TR-77-80

at a given frequency, the amount of time a stage (or group

of stages) will be in the highly loaded condition depends on

the amplitude of the imposed oscillation. Once a stage (or

group of stages) is transiently held in the adversely loaded

condition for a sufficient time period, the breakdown

process is locally like that which occurs in steady-state

loading.

The extent to which the flow coefficients oscil-

late depends on the dynamic response (phase angle and

amplitude ratio), which depends on oscillation frequency.

The loading process of Fig. 25 was repeated for frequencies

up to 320 Hz. The amplitude of imposed oscillation required

to drive the compressor to the stability limit is shown in

Fig. 26. As frequency increases, the amplitude required for

instability decreases. This occurs because, as frequency

increases, the phase lags and amplitude attenuations across

the stages increase; thus, the flow coefficient oscillations

are larger for a given imposed oscillation amplitude at

higher frequencies. (A fundamental discussion of the phase

shift and amplitude ratio effect is given in [24].)

A local minimum occurs in Fig. 26 near 80 Hz.

This occurs because, below 80 Hz, aft stages in the com-

pressor are critical, but, at frequencies higher than 80 Hz,

the amplitude attenuation across the forward stages

increases to the extent that it protects the aft stages, and

a new stage, farther forward in the compressor, becomes

critical.

68

A E D C-T R -77-80

The computations of Fig. 26 were made both with

and without the unsteady cascade correction to the stage

characteristics. First-stage reduced frequency is indicated

at each value of oscillation frequency. Only small

influences were experienced for this compressor in oscil-

latory inflow because of the relatively low reduced

frequencies. Compressors and fans with larger blades would

experience the influence of the correction at somewhat lower

frequencies.

No experimental data were available for this com-

pressor for direct comparison of stability limits. The

general trend with frequency and the amplitude magnitudes

are in agreement with the experiment of [37], however.

Oscillatin~ Inflow D~namic Response,

J85-13 Compressor

As indicated in the previous section, the ability

of the model to properly compute stability limits in oscil-

latory flow depends on its ability to compute the dynamic

response of the various stages in the compressor to the

imposed oscillations. Two experiments were carried out by

NASA using the J85-13 turbojet engine to measure the

unstalled dynamic response of the compressor to imposed

inflow oscillations [52, 53]. Figure 27 is a sketch of the

test arrangements. One test was conducted in a supersonic

wind tunnel with an axisymmetric mixed compression inlet

installed (Fig. 27a). Engine entry total pressure

69

AE DC-T R-77-80

oscillations were produced by oscillating the inlet bypass

doors, causing the terminal normal shock in the inlet to

oscillate, thus producing the downstream total pressure

oscillations. The other test arrangement (Fig. 27b) used an

airjet system in a section of ducting ahead of the engine.

The airjet system operates by blowing high velocity air

counter to the primary airstream. Planar oscillations in

engine entry pressure were produced by oscillating the air-

jet counterflow. In both cases, the oscillation amplitude

was approximately 10%, peak-to-peak.

The J85-13 compressor model was subjected to

imposed total pressure oscillations at the forward control

volume plane, thus simulating the inlet experiment of

Fig. 27a. The test setup of Fig. 27b differs from the model

and the inlet experiment, in that the forward isolation

plane provided by the supersonic flow in the throat of the

inlet is not in the airjet test setup. The inlet and model

dynamic systems consisted of the ducting downstream of the

normal shock wave to the choked turbine nozzle in the engine.

The airier dynamic system includes all of the test cell

ducting forward of the engine (until a choked valve,

venturi, or other forward isolation point is reached) down

to the choked turbine nozzle in the engine.

Figure 28 shows the variation of pressure amplitude

ratio and phase angle for a 40-Hz oscillation through the

compressor. Model computations are compared to the NASA

data. The calculated trends are in good agreement with

70

AE DC-TR-77-80

experiment. Normalized magnitude differs by approximatoly

10~ and phase angle by five degrees. The agreement,

although reasonable for dynamic response data, could be

improved by closer matching of engine operating condition

between model and experiment. Also, the J85-13 engine run

in the wind tunnel was likely not the same engine from which

the stage characteristics were taken; thus, engine-to-engine

variations are present. Further, the measurement uncertain-

ties present in the experimental data (quoted as ±3% on

pressure measurement and ±8 degrees on phase angle in [52])

may affect the comparison.

More detailed dynamic response results are shown

in Fig. 29. Figures 29a and 29b represent the frequency

response across one stage; Figs. 29c and 39d across approxi-

mately half of the compressor; and Figs. 29e and 29f across

the entire compressor. Trend and level agreement between

the model computations and the inlet results were reasonable,

but, once again, could be improved by closer matching of the

actual experimental hardware. In particular, the compressor

discharge agreement suffered because the combustor volume of

the actual engine was larger than that modeled. The com-

bustor in the model was shortened during steady-state

stability studies to keep the number of computation planes

to a minimum, and was not increased for the dynamic response

study.

Comparisons of the model and engine-with-inlet

results with the airier results show considerable

7]

AED C-TR-77-80

differences, especially near 40 Hz. The most likely cause

for the difference is the fact that the two dynamic systems

are not equivalent.

The model configuration used in the computations

was not tailored in detail to the experimental configuration

tested. The comparison shows that to obtain close agreement

dynamically, close attention to matching configuration must

be observed. The comparisons also show that considerable

care must be exercised in dynamic experimental work to

assure dynamic equivalency of systems (e.g., same isolation

points, volumes and geometry).

The results also suggest a potentially powerful

coupling of experimental work and analytical modeling. For

example, if it were not possible (or desired) to physically

dynamically isolate the airjet setup in the test cell to

simulate a given inlet-engine combination, the test setup,

including the airier system and the compressor, could be

modeled and the model refined using experimental results to

closely match the test setup measurements. Then, the

analytical model inlet configuration could be modified to

match any desired configuration, and the desired information

derived from the model.

Oscillatin@ Compressor Discharge Pressure,

J85-13 Compressor

Fuel flow to the J85-13 engine run with the super-

sonic inlet (Fig. 27a) was oscillated to determine the

72

AE DC-TR-77-80

propulsion system's dynamic response to oscillations induced

within the engine. The oscillations produced compressor

discharge oscillations of approximately 10%. The inlet was

assumed to remain started, thus, the model dynamic system

boundaries were the same as for the oscillating inlet bypass

door case. Computations were made using the J85-13 model by

adding heat to the combustor control volumes in an oscil-

latory manner to produce peak-to-peak oscillations in

compressor back pressure of approximately 10%.

Figure 30 provides a comparison between computed

and experimental amplitude ratio and phase angle variation

through the compressor at a frequency of 40 Hz. Trend and

level agreement were reasonable. Because of the high com-

pressor rotor inertia, at 40 Hz, rotor speed cannot respond

to the imposed fuel flow oscillations; thus, the effects of

the oscillation are from aerodynamic upstream propagation

only. The aft-stage characteristics (stages 4 through 8)

(Appendix B) are quite steep, and at this condition, the

stages are operating on the steep, negative slope part of

the curves. Therefore, perturbations in the aft-stage

airflow (or flow coefficient) will cause large oscillations

in stage pressure ratio. Figure 30 indicates the magnifi-

cation of the induced oscillation at the stage 4 entry.

However, stages i, 2, and 3 are operating in the flow coef-

ficient range of 0.5 to 0.55 (i.e., on the shallow, positive

slope part of the curves for stages 2 and 3, and the

shallow, negative slope range for stage 1). As a result,

73

AE DC-TR-7?-80

the forward stages oscillate pressure ratio very little in

response to a given airflow oscillation. Therefore, the

forward stages attenuate the upstream propagating dis-

turbance to the point that it is barely measurable at the

compressor inlet (Fig. 30a). Thus, it becomes clear that

the stage characteristic shapes and match points play a very

important role, not only in the steady-state stability of

the compressor, but also in the way disturbances are propa-

gated, thus, in the transient and dynamic stability

characteristics of the compressor.

Figures 31a and 31b present the dynamic response

to oscillating fuel flow across the compressor. Experi-

mental results are presented along with three different sets

of model computations. The first set is for constant rotor

speed and constant P1 and T I. The second set is for

variable rotor speed and constant P1 and T I. The third set

is for variable rotor speed, variable P1 and constant T I.

The three sets were computed to account for the three

dynamic effects caused by oscillating fuel flow. At fre-

quencies above approximately 20 Hz, the rotor speed cannot

respond to the oscillating fuel flow~ thus, as already

discussed, upstream aerodynamic propagation is the chief

mechanism present. At frequencies below 20 Hz, rotor speed

can respond to the oscillations in fuel flow. thus, the

stage flow coefficients must now vary mechanically also

(i.e., wheel speed oscillates). Further, because wheel

speed varies, compressor airflow demand also must vary.

74

AEDC-TR -77-80

Inlet airflow varies as a result and causes the normal shock

wave to oscillate. This causes the inlet total pressure,

PI' to oscillate.

Engine rotor speed response was represented in the

model by a simple first-order differential equation simu-

lating the engine control and rotor dynamic response. An

oscillation in Pl' at the control volume forward boundary,

was imposed to simulate the shock wave motion. As seen in

Fig. 31, reasonable agreement with experiments was obtained

only when all three of these effects were included. Because

of the complicated nature of the three disturbance paths,

agreement between model computations and experimental

results at internal compressor locations was poor. Agree-

ment could be improved by more exact modeling of the

specific engine under consideration.

This example makes clear the importance of

auxiliary perturbation propagation paths which must be con-

sidered, and which can have a strong influence on dynamic

response, thus on stability.

Rapid Inlet Temperature Ramp to Stall,

NACA-8 Compressor

The NACA-8 compressor model was subjected to a

rapid rise in inlet temperature simulating ingestion of hot

gases from gun or rocket fire. Mechanical rotor speed, N,

and P1 were held constant. The initial corrected rotor

speed was 70%. Figure 32a shows the imposed temperature

75

AE DC-TR-77-80

ramp. The engine entry temperature responded as shown in

the figure. Approximately 0.01 second was required for the

temperature pulse to be convected to the engine face.

Figure 32b shows the rapid reduction in corrected rotor

speed caused by the temperature ramp. Figure 32c shows the

stage flow coefficient variation caused by the temperature

ramp. The forward stages are rapidly driven downward by the

reduction in airflow accompanying the rapid temperature rise.

The resulting increase in forward stage loading immediately

prior to stall initially holds the mid-stage reduction to a

minimum, and actually unloads the aft stages. A point is

reached, however, where the forward stages stall and can no

longer support the load imposed on them. Flow breakdown

then occurs and is propagated back through the compressor.

In the example shown, the compressor may have

recovered frum the instability with the removal of the

temperature pulse. On an actual engine, combustor flameout

would probably have occurred as a result of the compressor

instability which was initiated.

A series of ramps of differing ramp height was

computed at a ramp rate of 3600"R/second. The results are

compared in Fig. 33 to a scatter band for three different

engines from [16]. Although Fig. 33 is far from an exact

comparison, the results do indicate proper trend and

magnitude.

76

AE DC-T R-77-80

Gas composition was considered constant for these

calculations. In the real experimental case, additional

effects may be present because of the gas composition

changes associated with ingestion of combustion products.

77

A E D C-T R -77-80

CHAPTER IV

DEVELOPMENT OF THREE-DIMENSIONAL MODEL

FOR DISTORTED FLOWS

The concepts used to develop the one-dimensional

model can be extended to allow computation of spatial dis-

tortion effects on stability. Figure 34 depicts the control

volume arrangement used for the distortion model. As was

noted in Chapter II, the one-sided finite difference form of

the governing equations could have been derived from a

finite control volume integral analysis rather than a

differential analysis. For greater simplicity in developing

the distortion model, the finite control volume, integral

approach is used.

The development of the distortion model is

described as follows. The governing equations are written.

Simplifications to allow distortion effects to be computed

with a minimum amount of internal radial and circumferential

flow detail are described. Special physical considerations

for radial distortion effects and for circumferential dis-

tortion effects are presented. The method of solution,

closely paralleling that used for the one-dimensional model

is then described.

78

A E D C -T R -77 -80

I. GOVERNING EQUATIONS

Figure 34 shows a finite control volume covering

an arc sector of a concentric ring of the compressor annulus

extending axially into the compressor one stage length. The

compressor is divided into control volumes as indicated in

Fig. 34. The subscript notation used to identify the

location of a control volume is also shown. Subscript "i"

indicates an axial position, "j" a radial position, and "k"

a circumferential position. Control volume i,j,k is there-

fore the control volume whose axial-facing entry plane is

on plane i, and whose axial-facing exit plane is on plane

ip = i + i. Its lower boundary surface approximately

facing the radial direction is on surface j, its upper face

is on surface jp = j + 1. Similarly, the control volume

surface normal to the circumferential direction is on

surface k and, advancing counterclockwise, its other surface

normal to the circumferential direction is on plane

kp = k + i. The compressor inlet is at i = i, the exit at

i = in. The annulus inner surface is at j = i, the outer

surface is at j = in. Circumferentially, k = i is the same

plane as k = kn + I. The compressor is therefore divided

into inm • jnm • kn control volumes where inm = (in-l) and

jnm = (in-l). For the example of Fig. 34, the compressor is

divided into 72 control volumes.

The compressor is divided radially on an equal

axially-facing area basis. That is,

79

A EDC-T R-77-80

AZtotal. = I (118) AZijk jnm • kn '

where AZtotal. is the total axial facing area at any plane, 1

i. Thus, all control volumes having the same index, i, are

of equal volume because their lengths are equal.

Figure 35 illustrates a single control volume.

For the purpose of this analysis, the control volume is

assumed approximately rectangular on each face and therefore

is a parallelepiped with the coordinate system; z, the axial

direction, r, the radial direction, and s, a coordinate

along the arc length of the control volume. The governing

equations for each control volume may be written as follows.

Mass

WZij k + WRij k + WCij k =

• i

mass flow entering the control v o l u m e p e r u n i t t i m e

WZipjk + WRijpk + WCijkp

v

mass flow leaving the control v o l u m e p e r u n i t t i m e

+ ~-~f cd(vol)

°lij k

Y

time rate of increase of mass " s t o r e d " i n t h e v o l u m e

(119)

80

AEDC-T R-77-80

where,

WZijk ' WRijk ' igk

are the mass flows into the axially-facing, radially-facing,

and circumferentially-facing surfaces, respectively.

Axial Momentum

FZij k + ~ + P S i j p k ARZijpk

+ PSij k AZij k -PSipjk AZipjk ]

v

axial forces acting on control volume

WZipjk Uipjk + WRijpk Uijpk + WCijkp Uijkp

v J

axial momentum leaving control volume per unit time

+

WZijk Uijk-WRij k Uij k-WCij k Uij k

t

axial momentum entering control volume per unit time

v (pU) d(vol) t olij k

'v

t i m e r a t e o f i n c r e a s e o f a x i a l moment~ 'n "stored" in the control volume

, . (120)

8!

AE DC-T R-77-80

The term, FZ, is the force of the blading and

cases acting on the fluid in the axial direction. In

Eq. (120), a projected area convention is used. The term,

ARZ, means the projected area in the axial (Z) direction of

the radially-facing (R) surface. S~milarly, the projected

area in the axial direction of the circumferentially-facing

area is ACZ, and is zero because of the way the control

volume was defined.

The bar above certain quantities of Eq. (120)

indicates that these terms actually represent surface

integrals. For example,

=~A PS d (ARZ) PSij k ARZij k

RZij k

(121)"

That is, the barred quantity indicates an integration of

pressure over the projected area of interest. l "

WRijpk Uijpk =I pvU d(ARR) .

~ARRijpk

Similarly,

(122)

Here, the barred quantity indicates integration over the

radial-facing surface area of the axial momentum being

carried across the surface by the radial flow. Implicit in

Eq. (120) is the assumption that the flow properties are all

uniform over the axially-facing surfaces of any given

control volume because no surface integration is indicated

over those surfaces.

82

AE D C-T R-77-80

Radial Momentum

[FRijk + PSij k ACRij k

+PSij k ARRij k

+ PSijkp ACRijkp

I -PSijpk ARRijpk ]

% ,%. •

radial forces acting on control volume

WZipjk Vipjk +WRijpk Vijpk +WCijkp Wijkp

• ~.. I

radial momentum leaving control volume

- WZij k vij k÷WRij k Vijk+WCij k Wijk

• v

radial momentum entering control volume

+ ~--~ (0v) d (vol)

Jvol...

time rate of increase of radial momentum "stored" in the control volume

. (123)

The force, FR, includes centrifugal forces and viscous

forces acting radially on the fluid in the control volume.

83

AE DC-TR-77-80

Circumferential Momentum

+

FCij k + ~ - P S i j k p ACCijkp

v

circumferential forces acting on c o n t r o l v o l u m e

WZipjk Wipjk +WRijpk Wijpk +WCijkp Wijkp

circumferential momentum leaving control volume per unit time

WZijk wij k-WRij k wij k-WCij k wij k

v

circumferential momentum entering control volume per unit time

f (ow)d(vol)

v°lij k

v

time rate of increase of circumferential momentum "stored" in the control volume

• (124)

The force, FC, includes Coriolis and viscous forces acting

on the fluid in the control volume.

84

A E D C -T R -7"/-80

~_n e.__rr g_.,y..

w

HZij k + HRij k +

v

enthalpy entering the control volume per unit time

+ WS + Q

v

shaft work done and heat added to fluid in control volume per

unit time

HZipjk + HRijpk + HCijkp

enthalpy leaving the control volume per unit time

(125)

+ 0f [ u2.v2..2] p e + 2 d(vol)

v°lij k

time rate of increase of energy "stored" in the control volume

In Eq. (125), the barred quantities represent integrations

over surfaces similar to those of the momentum equations.

For example,

HRijk = fAR [Cp TS +

U 2 + v 2 + w 2 pv d (AR) . (126)

8S

AE DC-TR-77-80

II. SIMPLIFICATIONS

In principle, the governing equations, Eqs. (119),

(120), (123), (124), and (125), could be solved directly if

all terms entering the equations could be evaluated. There

are five differential equations for each control volume;

therefore, for the example of Fig. 34, simultaneous solution

of 360 differential equations would be required. This, in

itself is not prohibitive, but, as finer divisions of the

compressor, or, as larger compressors are considered, the

computer storage and time requirements make the analysis

inconvenient to use. Also, direct solution of the radial

and circumferential momentum equations requires knowledge of

equivalent radial and circumferential body forces not

readily determinable. Further, numerical stability becomes

more and more of a problem as the level of complexity and

number of equations are increased.

The aim of this analysis is to compute the

influence of distorted flows on compressor stability. This

goal can be achieved with simplifications to the governing

equations which circumvent, at least to an extent, the

problems listed above and produce a practical model for dis-

tortion analysis. The simplifications are as follows.

Radial and Circumferential Flows

Each control volume of the compressor is assumed

to be representable as a compressor element which is in

radial and circumferential equilibrium with adjacent control

86

AE DC-T R-77-80

volumes when the compressor is running with uniform flow at

its entry plane and uniform throttling at the exit. This

means radial and circumferential crossflows are zero with no

distortion present. Further, it is assumed that with dis-

tortion present, the radial and circumferential flow

velocities are no greater than an order of magnitude less

than the axial velocity.

The circumferential and radial crossflows into the

control volume were computed using empirical relationships

in place of the circumferential and radial momentum

equations. Algebraic expressions, relating the crossflows

to the static pressure differences between volumes, were

used. This substitution reduced the number of differential

equations requiring solution from five times to three times

the number of control volumes.

Circumferential flows. Figure 36 illustrates a

pair of control volumes circumferentially adjacent but with

different static pressures. Looking down the blades in a

spanwise direction, the flow path for the circumferential

flow is through the blade row gaps. The blade row gaps, in

effect, form a series of orifice-like restrictions to the

circumferential flow. The circumferential flow is there-

fore computed by assuming that the static pressure in the

high pressure region, PShigh, is a "reservoir" pressure

driving flow to a lower pressure, PSIo w, across a series of

87

AE DC-TR-77-80

orifices. The sequence of equations comprising the approxi-

mating algorithm is as follows.

Ec I J PShi~h 7

I m I •

Mgap PSIo w (127)

Vgap = Mgap alow pressure ' (128) region

WCgap = Plow pressure Vgap Agap region effective

(129)

The approximations of using low pressure region acoustic

velocity and density are justified because the actual

differences between adjacent control volumes are not large.

The effective gap area is given by

A = CXFC A , (130) gap gap effective physical

where CXFC is the circumferential crossflow coefficient.

For most cases, a nominal value of CXFC = 0.6 was used. The

best way to determine the coefficient value would be to

empirically calibrate it using compressor test data to pro-

vide observed crossflow values with circumferential

distortion applied to the compressor. The physical gap area

used included the space between the rotor and the stator

internal to each control volume plus the gap between the

trailing edge of the stator upstream of the control volume

and the rotor of the control volume. For the first stage,

88

AE DC-TR-77-80

the gap area was doubled to provide an accounting for the

additional circumferential flow adjustment which occurs

upstream of the compressor first rotor. Experimental data

is given in [54] which indicates the extent of the cir-

cumferential flow redistribution occurring at the compressor

entry. The process is described in detail in [38]. The

impact of crossflow magnitude on the solution is indicated

in Chapter V.

The circumferential velocity component across the

face of a control volume was computed as

V +V gaPi~ k gaPijkp (131)

Wijk = 2

The averaging is performed because the "gap" values apply

at the circumferential boundaries of the control volume.

Radial flows. Figure 37 shows a pair of radially

adjacent control volumes with different static pressure.

Viewed looking axially into the compressor, the blades form

channels in which radial flow may occur. In the actual

compressor, strong secondary flow fields caused by rotation

and viscous effects exist [55]. A considerable amount of

mass is carried back and forth across the control volume

radial boundary on a scale usually smaller than, or on the

order of, the control volume size. For the purposes of this

analysis, it is the net flow exchange, caused by distortion,

89

A E D C-T R -77 -80

which is desired. An approximate, empirical form of the

radial momentum equation is used.

WRnet = CXFRVPlow [ PShigh-PSlow] +FR • (132)

The form of the equation is a simplification of the full

radial momentum equation produced by neglecting the time-

dependent term and flux terms across all but the radial-

facing control volume boundaries. The force term, FR, is

representative of the sum of viscous and centrifugal forces,

2 FR=Fviscou s +~ vol ~ r . (133)

Because WRne t is the radial crossflow caused by distortion,

the force, FR, is adjusted so as to cause WRne t to be zero

with the compressor running with no distortion. This is, in

a sense, equivalent to imposing the radial equilibrium

condition,

2 ~P p c, r ar = ' (134)

often used in three-dimensional analysis of compressor

flows [20]. (In this application, however, FR conceptually

includes viscous forces as well as centrifugal forces.) If

the stage characteristics of each radially adjacent control

volume were the same, with no distortion present, FR would

be zero. If radial variation of stage characteristics is

90

A E D C-T S -77-80

considered (treated later in this chapter), FR is set to

cause WRne t to be zero with no distortion present.

The term, CXFR, is the radial crossflow coef-

ficient, and was determined by subjecting the model to

radial inlet distortion and adjusting the coefficient to

permit the stage exit radial static pressure gradient to

vary as much as 2% from the undistorted value. This pro-

cedure was adopted because experimental data on radial

distortion indicate that although the induced static

pressure gradients relax as the flow progresses through the

compressor, a slight gradient does remain [56]. The upper

limit of CXFR was set by model numerical stability con-

siderations. A value of CXFR = 0.01 was used for most cases

computed in this study. For modeling a specific compressor,

the value of CXFR would best be determined through tests of

a compressor with radial distortion imposed and, adjustment

of the coefficient made to match the observed change in

stage exit flow distributions caused by the radial

distortion.

The radial velocity across the boundary is

computed from,

WRne t VR = (135)

Plow Achannel

9!

AE D C-T R-77-80

The radial velocity component across the face of the control

volume was computed as

VRij k + VRijpk Vijk = 2 " (136)

III. SPECIAL CONSIDERATIONS FOR

DISTORTION EFFECTS

Radial Distortion

Figure 38 illustrates the key physical processes

occurring when a radially distorted flow approaches a rotor

blade. On the left of Fig. 38, with uniform flow, con-

sidering the "design A" case, the tip and hub angles of

attack and stage characteristic operating points are stable,

near design conditions. On the right side of Fig. 38, a hub

radial distortion (low velocity or total pressure in the hub

region) case is shown. For case "A," the tip section

operates at a lower angle of attack and higher flow coef-

ficient while the hub becomes highly loaded, and, in case

"A," is even stalled.

In the "design "B" case, the tip section in uni-

form flow is highly loaded while the hub section is only

lightly loaded. Introduction of the radial distortion

pattern for the "B" case causes the stage to operate at a

better condition than was experienced with uniform flow.

This illustrates that the way a compressor responds to

radial distortion must depend on the spanwise work

92

AEDC-TR -77-80

distribution. Therefore, the one-dimensional stage

characteristic used up to this point will not permit compu-

tation of radial distortion effects. As noted in Fig. 38,

the distortion induced shift in radial work distribution, as

well as the distorted flow pattern itself, induces radial

crossflows.

The three effects which must be considered for

radial distortion analysis are then as follows.

Radial variation of compressor entry flow

properties. These are automatically treated by the model

because of the division of the compressor into control

volumes radially as well as circumferentially.

Radial variation of work and force distribution.

Stage characteristics may be specified for each radial

location in the same manner as a single, average stage

characteristic was specified for the one-dimensional

analysis. For the example of Fig. 34, each stage would have

a hub, a midspan, and a tip set of stage characteristics.

For the XC-I compressor used in example calcu-

lations in Chapter V, the compressor was divided radially

into five annuli (jn = 6). Stage characteristics for each

annulus were not available, but radial distributions of

total pressure and total temperature at each stage discharge

were given in [43] for one operating point at the corrected

speed setting analyzed. These profile data were used to

compute a correction to stage pressure ratio and stage

93

A E D C-T R -77-80

temperature ratio as a function of radius for each stage;

thus, in effect, producing radially variant stage charac-

teristics. A total temperature ratio correction factor,

CF T(r) = TR(r) TR ' (137)

avg

and a total pressure ratio correction factor,

CF P(r) = PR(r) (138) PRavg '

were determined for each stage from the profile data of

[43].

become The stage characteristics as functions of radius then

-1 ~t _1 U 2

~T(r) =~Tvg CF T+~'w-~h|[CFT-1]

and

~P (r) = ~P 7 avg cFP

(139)

(140)

Inherent in this correction is the assumption that the

pressure ratio and temperature ratio profiles, normalized by

their averages, do not vary greatly over the range of flow

coefficient of interest. Examination of data from other

compressors where profile data were available over a wider

range of operating conditions than used in this study showed

the profiles' shapes to be consistent (within ±5%, and

usually better) up to the point of stall.

94

AEDC-TR-77-80

Radial crossflows. Radial crossflows are computed

as described previously. A stage coefficient correction for

the distortion-induced radial flow velocity across the

control volume face is made by computing ~ as

U2 - v 2 ¢ U2 . (141)

wh

Circumferential distortion steady flow effects.

Figure 39a shows a compressor subjected to a simple circum-

ferential distortion pattern. Half of the face of the

compressor is in a lower than average total pressure region

and half is above. The example distortion waveform is a

simple sine wave. Upstream of the compressor face a length

of one compressor diameter or more, the static pressure

variatlon across the entry duct caused by the presence of

the compressor in the distorted flow field will be nearly

zero. The total pressure distortion at that location is

then principally an axial velocity distortion. Depending on

design characteristics and operating conditions of the

compressor, a portion of the high velocity region will be

diffused, creating a circumferential static pressure

gradient at the compressor face. This results in the

generation of a swirl velocity (Fig. 39a) carrying mass from

the high to the low pressure regions of the compressor face.

(This phenomenon is described in more detail in [27, 38,

54].) Figure 39a shows the influence of the variation in

95

A EDC-TR-77-80

axial and swirl velocities at four circumferential locations.

At 8 = 0, there is no swirl velocity and axial velocity is

high; therefore, ~ is low and the stage is operating at a

lightly loaded condition. At 8 = 90 °, the axial velocity is

still fairly high, and the swirl component exists, and is in

the direction of rotor rotation. The swirl component

combines with wheel speed to reduce the angle of attack at

this point. At 8 = 180 ° , the axial velocity is at a

minimum, thus causing a high angle of attack and high, even

stalled, stage loading. At 8 = 270 °, the axial velocity is

increased once more, but the swirl velocity component is

counter to rotor speed, thereby causing a higher angle of

attack and stage loading.

To a lesser extent, the same flow situation exists

at each stage entry. The effects on the stage character-

istic operating points are accounted for by computation of

the flow coefficient as

U = +V " (142)

Uwh sw

The swirl velocity, Vsw , is simply the circumferential

velocity component, w, at that particular location.

Unsteady effects. Figure 39b illustrates unsteady

airfoil effects for the same circumferential distortion

pattern. As indicated in Fig. 39b, the axial and swirl

velocities caused the flow coefficient to decrease, even to

96

AE DC-TR-77-80

below the steady-state stage stall value. The distortion

pattern is stationary. The rotor blades pass through the

spatially variant stationary pattern with each rotation,

and, from their point of reference, the flow is cyclically

unsteady. For a sinusoidal pattern, the frequency felt by a

rotor blade is

N f = 2--~ = ~ , (143)

where N is the mechanical rotor speed in rpm.

If the distortion pattern contained two cycles of

a sine wave instead of one (i.e., two low pressure regions),

the frequency would double. The reduced frequency for the

blade is

k = 2~fC _ NC (144) Vre I 60 Vre I

As discussed in Chapter II, the blade forces cannot instan-

taneously respond to rapid changes in angle of attack.

Therefore, as a blade passes through the low pressure region

(region of low %, Fig. 39b), the blade forces, thus ~, do

not follow the # variation in a steady-state manner. Near

the stall point, if the variation is sufficiently rapid, the

lag is such that ¢ may proceed into the stalled region, and

quickly retreat, without blade stall actually occurring

[35]. The effect is more pronounced as reduced frequency

increases.

97

AE DC-T R-77-80

The unsteady cascade aerodynamics arising Item

circumferentlal distortion are approximated in the model

following the treatment described in Chapter If, with an

empirical correction added for the stalled blade region of

operation. In Figs. Ii and 14, the unsteady lift magnitude

is given as a function of reduced frequency. As was shown

in Chapter II, Eq. (58), the dynamic response of ~ is the

same as the dynamic response of C£. Therefore, the ratio of

maximum dynamic ~ to the quasi-steady-state value is the

same as the corresponding lift coefficient ratio. Or,

I Imax Ic Imax Iss Ic Iss

= DLR . (145)

(DLR means dynamic lift ratio.)

Referring to the upper left-hand corner of

Fig. 39b, as a rotor blade begins at location k = I, the

value of ~ is ~i" As it progresses to the low pressure

area, k = 4, ~ = ~4" The value of #4 calculated on a

steady flow basis in region k = 4 is ~4 , but, is incorrect ss

because of dynamic effects. Let the actual value be ~4dy n.

A correction may be constructed as follows.

- 41 I ¢ I ma x ~4d~ n - = DLR (146)

ss ss

98

A E D C-T R -77-80

Therefore,

4dy n ' + (i - DLR) ~i " = DLR ~4s s (147)

Thus, the actual ~4 value depends not only on the steady-

state value computed at the k = 4 location, but also on the

blade's history as it passed from location k = I.

This dynamic lag algorithm can be more generally

written so as to be applied successively from one sector to

the proceeding sector in the direction of rotor rotation.

Or,

~ijk = DLR ~ijks s + (i - DLR) ~ijkm " (148)

Thus, a lag is introduced into the stage characteristics.

The value of DLR used was computed from

i

DLR = -~/ 1

V 1 + (ak) , (149)

as shown in Fig. 14. The reduced frequency, k, depended on

the shape of the circumferential distortion pattern. The

value of £ used followed the Goethert-Reddy analysis, but

could readily be varied based on measured response of a

compressor to circumferential distortion of varying pattern

shape to account for additional viscous effects. This

part of the unsteady correction predominates in the blade

unstalled region of operation.

99

AE DC-TR-77-80

Goethert and Reddy point out in [32] that blade

dynamic response should be expected to be more sluggish

because of boundary layer effects not included in their

inviscid analysis. This will be particularly true when the

blade is operating in its stalled region. As already

pointed out, the blade may transiently operate at angles of

attack greater than the steady-state stall angle of attack

without actually stalling. Experimental results on the

degree to which this phenomenon may be present for isolated

airfoils were compiled by Melick [36] and are indicated in

Fig. 39c. The dynamic stall lift coefficient increment is

correlated to the nondimensional time derivative of angle of

attack for several different types of isolated airfoils.

This effect was incorporated in the model by

computing the derivative of angle of attack from the change

in flow coefficient as a blade passed through the distortion

pattern using Eq. (34) to relate flow coefficient to angle

of attack. A stall lift coefficient increment was then

calculated using the empirical relationship of Fig. 39c.

The resulting lift coefficient increment was then used to

compute an equivalent increment in stage pressure and

temperature coefficients using Eq. (39), neglecting the

influence of the blade drag coefficient. With this infor-

mation, the stage characteristic curve was translated

upward, at the unstalled stage characteristic slope existing

near the stage characteristic peak, in a manner analogous to

the CE - ~ curve translation of Fig. 39c. This translation,

100

AEDC-TR -77-80

when combined with the dynamic lag effects of Eq. (149)

yields the dynamic stage characteristic behavior in the

stalled stage region depicted in Fig. 39b.

The Goethert-Reddy analysis [32] and Fig. ii

indicated less dynamic effects with a cascade of airfoils

than are experienced with a single, isolated airfoil. It

might well be anticipated that the stall region behavior of

the cascade would be less than the isolated airfoil case as

well. For that reason, the "stall overshoot" constant, m,

Fig. 39c, was chosen at four different values, and influence

on the solution determined parametrically. A value of m

between one and ten was found to be reasonable. Details of

the parametric investigation are discussed in Chapter V.

IV. METHOD OF SOLUTION

In this section, the approximation to the time

derivative of the volume integrals in the governing

equations is described. The resulting model equations,

including effects of the simplifications already described,

are written and the solution procedure described.

Volume Integral Time-Derivative Approximation

In each of the governing equations, Eqs. (119),

(120), and (125), a partial time derivative of a volume

integral exists. The general form is

101

AEDC-TR-7740

Y = ~t l~" y (z,r,s,t) d(vol)

#v ol

= f ~Y (z,r,s t) d(vol) ~t ' " Jv ol

(150)

It is desired to approximate Y in terms of a control volume

surface value of y. By the mean value theorem, there exists

some point (z,r,s) interior to the control volume, such that

fro ~ (z'r's't) d(vol) = (vol} ~ (z,r,s,t) .

1

(151)

~Yo Let -~£- be a value of the integrand on either the forward or

aft control volume face normal to the z-axis. The time

derivative at the mean value point is then related to the

surface time derivative by expanding around the surface

point,

where

~Y (z,r s t)=°Y [ z O So t ] ~t ' ' ~t 'ro' '

Z O +a~Z,ro,So,t ] ] Az

[ "o,ro, o

o < a, b, c < 1

, (152)

102

AE DC-TR-77-80

Also,

and

AZ = Z - ~ O

Ar= r - r o

AS = S O -

(153)

The approximation,

~Yo ¥ = (vol) ~- , (154)

is chosen for use in the model. Multiplying Eq. (152) by

the volume, and rearranging the order of differentiation

yields an estimate for the truncation error.

E -" (vol) ~-~ AZ + ~r

The spatial derivatives are finite; therefore, as

Az , Ar , As ~ 0 ,

(155)

the error, c, vanishes, thus the approximation is consistent

of order one.

Numerical stability was investigated by running

with different time step sizes as in the one-dimensional

analysis. A time step size of approximately one-half the

value indicated by the Courant-Friedrichs-Lewy criterion

was used.

103

AE DC-TA-77-80

The actual approximations made were,

Mass

p d(vol) = vol ~ Pipjk '

d volijk

(156)

Axial Momentum

~-~ (p U) d(vol) = vol (p U)

v°lij k

ijk ' (157)

Energy

~fv Pie+

°lij k

U 2 + v 2 + w 2 2 J d (vol)

= vol ~ p e+ 2 ipjk

(158)

Model Equations

Using the approximations and simplifications pre-

viously described yields the equations comprising the

distortion model.

Governing Equations

Mass, axial momentum, and energy were used. The

assumption of small v and w permits ignoring v and w terms

104

AEDC-TR-77-80

in some instances. Also, radial and circumferential flows

were determined by the crossflow approximations, thus

replacing the radial and circumferential momentum equations.

The resulting equations are as follows.

Mass

a ipjk 1 [ ~t - volij k WZijk + WRijk + WCijk

- WZipjk -WRijpk - WCijkp ] " (159)

Axial Momentum

8 ( P U)i~k = 1 [ ] (160) ~t ~ Fijk + IMPijk - IMPipjk "

In Eq. (160) the contribution of the pressure integrated

over the projected areas of the radially oriented surfaces

is combined with the force term, FZ, of Eq. (120) to pro-

duce F. The force, F, is obtained from stage characteristic

information, as was done in the one-dimensional case. The

impulse terms are

IMP = PS AZ + WZ U . (161)

The momentum convected across the radial and circumferential

boundaries is negligible compared to the axial impulse terms.

105

AED C-TR-77-80

Energy

xi [ ] -- 1

(vol)ii k ~ijk + QiJk + I~ijk - HZipjk • (162)

In Eq. (162),

[ u2 ] X = p e + -~- . (163)

The v 2 and w 2 terms in Eq. (125) are negligible compared to

U 2 .

The terms neglected in Eqs. (160) and (162) are

second-order terms determined by nondimensionalization and

order of magnitude analysis of the equations• The terms

were ignored because their influence on the solution was

small. Also, inclusion of terms caused the numerical solu-

tions to be generally less stable and to oscillate with

small amplitude when a steady-state condition was reached.

Solution Procedure

The solution procedure used on the digital

computer paralleled the approach used for the one-

dimensional model outlined in Fig. 16. Initial values of

the dependent variables, D, P U, and X, were specified for

every control volume location, ijk, to start the solution.

For simplicity, the solution was always started from a

steady, uniform flow condition, thus easing the chore of

computing initial conditions.

106

AEDC-TR-77-80

Forward boundary conditions. Forward boundary

conditions were total pressure and total temperature,

Pl,j,k and Tl,j,k, specified at each control voiume face.

Entry plane conditions were then computed using the same

relationships as indicated in Fig. 16, but for every control

volume face instead of just one. Thus, the subscript, i, in

Fig. 16, entry plane conditions, is replaced with subscript,

l,j,k.

Interior plane values. The same relationships

indicated in Fig. 16 are used again. Once more, the

subscript, i, must be replaced with i,j,k, and the quanti-

ties computed at each control volume face.

Crossflow values. Crossflows among the control

volumes are computed using Eqs. (127) through (136) for each

control volume. For all calculations made in this study,

crossflow across the compressor walls was zero.

Exit plane conditions. Only unchoked flow at the

exit was used for this study; thus, a uniform static

pressure was specified at every control volume aft face at

the compressor exit plane. Exit plane calculations follow-

ing the unchoked portion of the one-dimensional procedure

(Fig. 16, exit plane conditions) were therefore used with

the index, in, replaced by in,j,k, and computations made for

all control volumes.

107

A E D C-T R-77 -80

Force and work computations. The same procedure

used in the one-dimensional procedure was used again, but

individually at each control volume. Local values of flow

coefficient were computed and adjusted for distortion

effects according to Eqs. (141) and (142). Further cor-

rections to stage characteristics for radial variations and

for unsteady effects were also made.

Time-derivative computations. The derivatives of

the dependent variables, D, p U, and X, are computed from

Eqs. (159), (160), and (162) for every control volume. The

values are then integrated to provide values of the vari-

ables at the next time step. The sequence is then repeated,

with the boundary conditions changing in accordance with the

specified problem under investigation and the computations

continued until the desired solution is obtained.

Additional numerical stability considerations. To

prevent numerical instabilities in the model solution caused

by error buildup in the crossflow calculations when the

driving pressure differences from control volume to control

volume are very small, the pressure differences are tested

to assure a significant pressure gradient exists. If a

significant difference exists, crossflow is calculated; if

not, crossflow is set to zero.

108

AE D C-TR-77-80

CHAPTER V

EXAMPLE PROBLEMS WITH COMPARISON TO EXPERIMENTS,

DISTORTION MODEL

One steady-state combined radial and circumfer-

ential distortion case for the XC-I compressor is given.

Detailed experimental data were available for this case and

comparisons between model computations and experimental

measurements are presented. An example of time-variant

distortion for the XC-I compressor is given. Pure radial

and pure circumferential distortion effects are computed.

A pure circumferential temperature distortion case is also

presented. Specific experimental data were not available

for the time-variant distortion, pure radial and pure

circumferential distortion patterns for the XC-I compressor.

Data from other compressor tests were used, therefore, to

check the model results for proper trend and order of

magnitude.

In all cases discussed in this chapter, the

nominal compressor entry conditions were one atmosphere

pressure and 518.7°R temperature.

I. COMBINED DISTORTION PATTERN, XC-1 COMPRESSOR

Figure 40 shows a contour map of the combined

circumferential and radial total pressure distortion pattern

to which the XC-I compressor was subjected. The pattern was

109

AED C-TR-77-80

generated using a distortion screen to simulate a distortion

pattern present in the inlet of a VTOL aircraft [43]. The

pattern of Fig. 40 is for a corrected rotor speed of 90%.

The pattern amplitude was 10.2%. The lowest pressure region

was centered at approximately 210 degrees, in the tip region.

The distortion pattern was input to the model by

dividing the compressor annulus as indicated in Fig. 414

Each stage was divided into 30 control volumes. The total

pressure at the compressor entry plane was averaged over

each control volume face to obtain the values indicated in

Fig. 41. The averaging resulted in reduction in the

apparent distortion pattern amplitude of approximately 1.0%.

The compressor was divided axially into four stages (five

axial locations) as shown in Fig. 34. Inlet total tempera-

ture was constant.

The compressor discharge static pressure was

specified as the aft boundary condition and was ramped

upward to load the compressor to the stability limit. The

process was conducted with and without distortion. The

results are shown in Fig. 42, and are compared to experi-

mental measurements. A reduction of 5.3% in stability limit

pressure ratio caused by distortion was computed. The

results agreed well with experimental measurements.

Figure 43a presents a detailed comparison of the

circumferential total pressure profiles in the tip region of

the compressor at each stage entry and exit. The distortion

pattern is seen to attenuate until, at the compressor

110

AEDC-TR-77-80

discharge, the variation is within 1% of the average value.

The agreement between the model and experiment is reasonably

good, usually within i%, but occasionally different by as

much as 2%.

Two mechanisms serve to attenuate the circumfer-

ential distortion. The first stems from the fact that in

the low pressure region, the flow coefficient is lower;

thus, if the stage is not stalled, it will work harder

(produce a higher pressure and temperature coefficient) than

the high pressure regions. This effect tends to drive the

stage discharge pressures of the low and high pressure

regions toward equality. Figure 43b shows the corresponding

total temperature profiles and indicates the transposition

of the pressure distortion to temperature distortion. The

second mechanism working to smooth out the circumferential

pressure profile is circumferential crossflow, which was

discussed in Chapter IV. The slight phase shift between the

experimental and computed profiles may be caused in part by

slightly different crossflow in the actual compressor than

that allowed for in the model. (Crossflow effects are dis-

cussed in more detail later in this chapter.)

Another reason for the apparent phase shift is the

finite size of the control volumes. A finer circumferential

division would reduce any skew of the computed profile

waveform caused by averaging. The profile waveform ampli-

tude is also reduced by the averaging effect of the finite

control volume, just as occurred between Figs. 40 and 41

111

AE DC-TR-77-80

Thus, the model profile amplitudes should be expected to be

1% or so less than the experimentally observed profiles.

Figure 44 shows the attenuation of the distortion

pattern through the compressor for the tip (Fig. 44a),

mid-span (Fig. 44b), and hub (Fig. 44c) regions. Agreement

between the computed and experimental results is generally

within 1% when allowance is made for the averaging effects

of the finite control volumes of the model. Pressure dis-

tortion reduction and tomperature distortion production are

similar for the tip and mid-span regions. In the hub

region, however, both the computed and experimental measure-

ments indicate amplification of the distortion across the

first stage. This occurs because of the shallower stage

characteristics which exist in the hub region. The agree-

ment is not as good in the hub region as in the tip and

mid-span regions, indicating some inaccuracy in the

estimation of the hub region stage characteristics.

Figure 45 presents a comparison of radial total

pressure and temperature profiles in the region of maximum

radial distortzon (at 210 degrees). Figure 45a shows agree-

ment within 1 to 2% except in the hub region at the second

stage entry. Once more, this indicates some inaccuracy in

the hub region stage characteristic. Figure 45b shows the

total temperature radial profiles. Agreement within 1% was

generally achieved.

Figure 46 presents similar profiles for the

compressor operating without distortion. Agreement between

!12

A ED C-TR-77-80

computed and experimental results was generally within 1%.

Comparing Figs. 45 and 46 indicates that the radial dis-

tortion with low pressure in the hub decays rapidly in the

compressor, but some modification to the profiles does

occur.

Computation of proper distributions within the

compressor is important to stability limit determination

because it determines the local stage flow coefficient

distribution which in turn controls stability. Figure 47

presents the approximate flow coefficient distribution

computed for each stage just prior to reaching the com-

pressor stability limit. The flow coefficient values are

normalized by the stage stall value. As should be expected,

the distribution approximates the shape of the total

pressure distortion pattern• Stage i, Fig. 47a, is almost

stalled in a small region at the tip. Stage 2, Fig. 47b, is

actually operating with the tip stalled in the distorted

region. Figure 47c shows a similar pattern for stage 3, but

it is not stalled. Stage 4, Fig. 47d, is stalled over a

wide arc from hub to tip, and, along with stage 2, is indi-

cated by the model to cause the flow breakdown leading to

total compressor instability.

The stalled region in stage 4 is seen to extend

beyond the concentrated low pressure region of the imposed

distortion pattern. The "pattern spreading" is caused by

113

AED C-TR-77-80

the attenuation mechanisms discussed previously. When a

sufficiently large region becomes affected, total insta-

bility will occur.

Control Volume Size Selection

The example problem discussed in this section

illustrates the division of the compressor into finite

control volumes. The circumferential division (60-degree

arc sectors) was made because, as Goethert and Reddy point

out in [32], it is at approximately 60 degrees that each arc

sector acts very much as a separate compressor (as parallel

compressor theory would suggest); thus, the effects of the

unsteady cascade aerodynamic and crossflow approximations

are less crucial to achieving an accurate representation of

the compressor with regard to circumferential distortion.

As already noted, some error is introduced into the model

results because of local distortion pattern variations

within a 60-degree arc sector which are averaged out by the

approximating process.

The radial division selection (five rings) was

made based on the number of radial locations at which

experimental total temperature and total pressure profile

data were available for approximation of the radial vari-

ation of stage characteristics. In general, the selection

of the radial division must depend on the strength of the

radial gradients of the distortion pattern and the internal

radial profiles within the compressor. Divisions which

I14

AE DC-TR-77-80

hold averaging error to within 1 or 2% should be quite

sufficient. Also, the radial crossflow representation in

the model is relatively crude. Therefore, it should not be

expected that finer and finer radial division of the com-

pressor would yield greater and greater accuracy of the

computed results. Additionally, the stage characteristic

data are based on one-dimensional (or, at best, averaged

over a radial distance) data, thus, a very fine division

radially would not be consistent with the resolution of the

input empirical information.

II. TIME-VARIANT DISTORTION

Figure 48 illustrates time-variant distortion.

The total pressure at the engine entry plane fluctuates with

time and also varies spatially across the compressor face.

The fluctuations are caused by unsteady flow processes in

the inlet, such as unsteady shock wave-boundary layer

separation [6, 13, 15]. The unsteady, nonuniform flow

causes distortlon patterns at the engine entry which vary

with time, as illustrated in Fig. 48. The spatially

averaged pressure also oscillates.

The XC-I compressor model was subjected to tlme-

varlant distortion by superimposing a time-dependent

waveform on the distortion pattern of Fig. 40. Figure 49

shows the frequency spectrum of the superimposed fluctu-

ations. The spectrum is representative of those measured

experimentally in aircraft inlets [15]. The waveform was

115

A E DC-TR-77-80

applied in two parts. One part was phase shifted over the

compressor face so as to alternately stretch and shrink the

steady-state distortion pattern. The other part was added

uniformly to produce timewise fluctuation in the spatial

average. (Details of time-variant distortion pattern

analysis and synthesis are discussed in [42].)

Two cases were computed for the XC-I compressor

with time-variant distortion. One was for an overall root-

mean-square (rms) amplitude of 2% and the other was for 4%.

With these inlet flow fields present, the compressor exit

pressure was ramped upward until instability occured.

Figure 50a shows the loss in stability limit pressure ratio

correlated with the rms amplitude. A reduction results, as

should be expected, and is of similar magnitude as the

experimental results of [7] for a different compressor.

The model computed results were also correlated to

the amplitude of the instantaneous distortion pattern

existing just prior to compressor instability, as shown in

Fig. 50b. All pattern shapes were similar, thereby per-

mitring use of the simple distortion pattern magnitude

descriptor. The results generalize well with steady-state

distortion results, thus providing a degree of analytical

conformation to the quasi-steady-state distortion approach

currently used in experimental stability analysis. (See

Chapter I and [7] for further details on that approach.)

The results shown in Fig. 50 should not be sur-

prising because the XC-1 compressor is very small (about

116

A E D C-T R -77-80

17 inches in diameter and 8 inches long) and has high rotor

speed (about 16,230 rpm at 90% speed). Therefore, fluctu-

ations in the frequency range considered are sufficiently

low to permit the instantaneous distortion pattern to dwell

at the entry, and penetrate through the compressor in a

quasi-steady manner. For example, in this case, approxi-

mately 0.0037 second is required for one rotor revolution,

and 0.0017 second is required to convect a particle through

the compressor (average axial velocity of 400 ft/sec). The

period of the highest frequency component in the spectrum is

0.0038 second. Thus, the distortion effects will be much

like steady-state distortion.

If larger amplitude timewise fluctuations are con-

sidered, particularly if the compressor is combined with

upstream ducting and a combustor, planar oscillation effects

similar to those discussed in Chapter III would show an

effect also.

III. PURE RADIAL AND CIRCUMFERENTIAL

DISTORTION PATTERNS

Three additional types of distortion patterns were

computed to gain some insight into the influence of pattern

severity and shape v~riations. The patterns considered are

pure radial pressure distortion, pure circumferential

pressure distortion, and pure circumferential temperature dis-

tortion. No specific experimental data for the XC-I

117

AE DC-TR-77-80

compressor were available for these cases. Comparisons to

experimental results from other compressors are made

therefore.

Radial Pressure Distortion

Figure 51a illustrates the results of loading the

XC-I compressor model to the stability limit with tip radial

distortion (low pressure in the tip region). Pattern

severities of 5, i0, and 20% were considered. Because of

the relatively high tip loading of this compressor, Fig. 46,

it appears to be more sensitive to tip radial distortion.

Figure 51b shows the results for the hub radial case and

indicates the compressor to be more tolerant. Figure 52

shows comparisons of the XC-1 model results with experi-

mental data from a J85-13 test [16]. Similar results were

obtained, except that the J85-13 was even more tolerant of

hub radial distortion than the XC-I was computed to be.

Circumferential Pressure Distortion

Figure 53 shows the effect of increasing the

severity of a 60-degree-arc circumferential distortion

pattern. Pattern magnitudes of 5, i0, 15, and 20% were

computed. Increasing loss of stability limit pressure ratio

and airflow was computed. Figure 54 shows the model

results compared to a range of experimental data from other

compressors. The model results appear reasonable.

Figure 55a depicts the variation of airflow in the

distorted region. Crossflow into the distorted region

118

AEDC-TR-77-80

rapidly increases the flow in that sector. Also shown in

Fig. 55a is the crossflow into an approximately 90-degree

distorted region computed for the XC-I compressor from

experimental measurements [43].

Two different crossflow distribution cases were

computed to check the influence of crossflow on computed

compressor behavior. In one case, the first stage cross-

flow was increased. The increase was accomplished by

increasing the first stage crossflow area. Figure 55b indi-

cates the effect of the crossflow change on compressor

stability. The increased first stage crossflow case

matched the experimental distribution better, and caused a

higher stability limit pressure ratio with distortion

present. The improved performance comes about by allowing

crossflow to relieve downstream stage loading in the dis-

torted region. The crossflow distribution more closely

matching the experimental case was used in the previously

discussed combined distortion case.

Figure 56 shows the influence on stability of

varying the angular extent of the circumferential low

pressure area while holding the distortion pattern magnitude

constant. Three different cases, 6- = 30, 60, and 180

degrees are shown. The 30-degree case has the least loss in

stability. The 60- and 180-degree cases have approximately

the same percentage loss in stability limit pressure ratio

(taken at constant corrected airflow through each stability

limit point). However, the flow and pressure ratio

119

AE DC-TR-77-80

reduction for the 180-degree case was more than that of the

60-degree case. Figure 57 shows the loss in stability limit

pressure ratio as a function of distortion pattern angular

extent. Computations were made for four different values of

dynamic lift coefficient "stall overshoot" constant, m,

Fig. 39c. Also shown in Fig. 57 are experimental data from

a J85-13 compressor subjected to distortion of varying

angular extent. The experimental results fall between the

m = 1 and m = i0 model computation results. The m = 40

results, corresponding to the center of the isolated airfoil

correlation curve, Fig. 39c, produce virtual insensitivity

to circumferential distortion up to 60 degrees, which is an

unrealistic result. A value of m between 1 and i0 is a

more realistic value. The results shown in Fig. 56 are for

m = i0.

The analysis presented in Fig. 57 indicates that

the unsteady cascade aerodynamic effects predominate in

determining the response of the compressor to circumfer-

ential distortion below approximately 60 degrees angular

extent.

Circumferential Temperature Distortion

Figure 58 shows the effect of varying the severity

of total temperature distortion on compressor stability.

Temperature distortion destabilizes the compressor in much

the same manner as does pressure distortion. The distorted

region, which is hotter than the rest of the compressor,

120

AEDC-TR-7?-80

operates at reduced airflow, thus at reduced flow coef-

ficients compared to the rest of the compressor. Therefore,

as the load on the compressor is increased, the hot region

stalls first, eventually causing the entire compressor to

become unstable.

Mechanical rotor speed was constant during the

loading. Corrected rotor speed, indicated in Fig. 58,

decreased as much as 1.1% because of the increased average

temperature over the face of the compressor.

Little influence of crossflow was found, indi-

cating that basic parallel compressor theory (with

appropriate unsteady cascade corrections) should give

reasonable predictions for temperature distortion.

Figure 59 shows a comparison of model computed

stability limit pressure ratio reduction to experimental

results from other compressors. The predicted trend is

reasonable, although more pressure ratio reduction is

computed than was experienced on the other compressors.

121

AEDC-TR-77-80

CI~PTER VI

SUMMARY OF RESULTS AND RECOMMENDATIONS

FOR FURTHER WORK

The overall objective of this work was to develop

an improved analytical method for the computation of turbine

engine compressor stability loss caused by a wide range of

different types of external flow disturbances. Another

objective was to gain understanding of the physical

processes leading up to instability as induced by external

disturbances. The understanding was provided by the

development of an analytical model and by application of the

model to example problems with comparisons to experimental

results. The objectives were accomplished in two major

steps. In the first step, a one-dimensional, time-dependent

math model was developed and applied to a variety of steady

and time-dependent planar disturbances which are destabi-

lizing to the compressor. In the second step, the concepts

of the one-dimensional model were extended to three

dimensions to allow the development of a time-dependent

model for computation of radial, circumferential and

combined distortion effects on stability. The distortion

model was then applied to the computation of distortion

cases with comparisons to experimental results. The results

of those two steps are summarized as follows.

122

AEDC-TR-77-80

I. ONE-DIMENSIONAL, TIME-DEPENDENT ANALYSIS

Model Description

The one-dimensional model satisfies the time-

dependent mass, momentum and energy equations. The full,

nonlinear form of the equations, not used in earlier works,

was used. The compressor is divided into control volumes on

a stage-by-stage basis. The force and shaft work applied to

the fluid is determined from empirical steady-state stage

characteristics modified for unsteady cascade airfoil

effects. The solution of the model governing equations is

obtained through finite difference approximation of the

governing differential equations. The resulting system of

equations was solved on a digital computer.

Example Problems

Steady-state stabilit~ limit. Three different

compressors were modeled. They were, an Allison XC-I lift

engine four-stage compressor, a NACA eight-stage research

compressor, and the General Electric J85-13 turbojet engine

compressor. The model was shown to be able to generally

compute the steady-state stability limit pressure ratio and

corrected airflow to within 1 to 2% of experimentally

determined values. Compressor internal steady-state flow

conditions in agreement with experimental values were also

shown to be computed by the model. The initial flow

123

AE DC-TR-77-80

breakdown process in the compressor at the stability limit

was shown to be reasonably well reproduced by the model.

D~namic stabilit~ limits (oscillatin~ inflow

loadin~)~ NACA ei~ht-sta~e compressor. The NACA eight-stage

compressor model was driven to its stability limit by planar

oscillation of the compressor inlet pressure at frequencies

from 20 to 320 Hz. The model results indicated that, as

frequency increased, the amplitude required to cause insta-

bility decreased. It also showed that the individual stages

transiently enter and retreat from their stalled regions

during the oscillations without total compressor instability

occurring. As the oscillation amplitude increased, however,

the time period during which the stage is in the stalled

region is increased, and a point is eventually reached where

total compressor instability occurs. The analysis showed

further that, for planar oscillations in the frequency range

analyzed, the influence on stability of the unsteady cascade

correction to the steady-state stage characteristics was

small.

The ability of the model to compute dynamic

stability limits was shown to depend strongly on the model's

ability to properly compute the dynamic response of the

compressor to imposed disturbances.

Oscillating inflow dynamic response, J85-13

compressor. The unstalled, dynamic response to planar entry

pressure oscillations in terms of amplitude ratio and phase

1:24

AE DC-TR-77-80

angle were computed and compared to experimental data. The

experimental data were obtained by NASA in a wind tunnel at

Math 2.5 with the engine operating behind a supersonic

inlet. The compressor entry pressure oscillations were pro-

duced by forced oscillations of the inlet bypass doors.

Data were acquired from 1 to 50 Hz. Amplitude ratios and

phase angles computed by the model agreed with experimental

values to within 10% and five degrees, respectively. An

additional set of experimental oscillating entry pressure

data obtained by NASA was compared to the inlet data and

model computations. The additional set was obtained in a

test cell with an array of small airjets blowing against the

primary engine flow in an oscillatory manner to produce the

pulsations. The results differed markedly from the inlet-

engine results, probably because of differences in the

boundaries of the dynamic system under consideration.

Oscillating compressor dischar@e pressure, J85-13

compressor. The unstalled, dynamic response of the J85-13

compressor to oscillations in discharge pressure caused by

engine fuel flow oscillations was computed using the model

and compared to experimental results. A frequency range of

i to 60 Hz was considered. Agreement between model compu-

tations and experimental results was 10 to 20% and 5 to i0

degrees on amplitude ratio and phase angle, respectively, at

40 Hz. The experiment was conducted in a wind tunnel at

Mach 2.5 with the engine installed behind a supersonic inlet.

125

AEDC-TR-77-80

Three dynamic disturbance paths were discovered to

exist at frequencies below 20 Hz and were included in the

model analysis. The three paths are: (I) the direct up-

stream propagation of the pressure disturbances, (2) oscil-

lation of stage pumping because of rotor speed oscillations,

and (3] oscillation of compressor entry pressure caused by

the supersonic inlet shock wave oscillations resulting from

the engine airflow oscillation. The example illustrated the

influence of multiple disturbance paths on dynamic behavior

and stability.

Rapid inlet temperature ramp, NACA eight-stave

compressor. The NACA eight-stage compressor model was

driven to instability by ramping inlet temperature at a rate

of 3600OR per second to simulate hot gas ingestion from

rocket or gun fire. Ramp heights of 100°R or over caused

instability. The results were compared to a band of experi-

mental data from other compressors. The model computations

were within the band.

II. DISTORTION ANALYSIS

Model Description

The concepts of the one-dimensional model were

extended to allow computation of the effects on stability of

distorted flow at the compressor entry plane. The com-

pressor was divided radially and circumferentially as well

as axially into control volumes, and crossflow between the

126

AEDC-TR-77-80

control volumes caused by distortion was approximated. The

governing equations were developed from the three-

dimensional, time-dependent mass, momentum and energy equa-

tions for a finite control volume. Empirical crossflow

relationships were used to replace the radial and circumfer-

ential momentum equations for purposes of simplification.

The radial work distribution variation effect necessary for

treatment of radial distortion was approximated using

average stage characteristics modified as a function of

radius by an empirical total pressure and total temperature

profile correction. Crossflow and unsteady cascade effects

for circumferential distortion were built into the model

also. The resulting set of equations was solved using a

digital computer.

Example Problems

Steady-state combined radial and circumferential

distortion. The four-stage, XC-I compressor was modeled for

distortion by dividing each stage into five radial and six

circumferential divisions. This produced 30 control volumes

per stage, or 120 control volumes for the entire compressor.

The model was subjected to a combined distortion pattern and

loaded to the stability limit using compressor discharge

static pressure as the aft boundary condition. Detailed

experimental data were available for this case and compari-

sons were made. The stability limit pressure ratio and

corrected airflow were computed to within 1.3 to 0.8% of the

127

AE DC-TR-77-80

experimental values, respectively. Detailed circumferential

and radial profile comparisons of total pressure and total

temperature were made. Agreement with experimental results

was reasonable.

Time-variant distortion. A timewise fluctuating

total pressure was superimposed on the steady-state combined

radial and circumferential distortion pattern to which the

XC-I compressor model was subjected to simulate time-variant

distortion. The reduction in stability limit pressure ratio

was shown to correlate well with the severity of the instan-

taneous distortion pattern existing just prior to compressor

instability. The correlation also agreed with results

obtained by imposing steady-state distortion of increased

severity on the model. This result provides an analytical

confirmation of the quasi-steady-state distortion method of

analysis currently used in experimental stability test and

analysis.

Pure Radial and Circumferential Distortion

The XC-I compressor model was subjected to pure

radial pressure distortion, pure circumferential pressure

distortion, and pure circumferential temperature distortion

patterns of varying severity and their influences on

stability were computed. No specific experimental data for

the XC-I compressor subjected to these specific distortion

patterns were available. Therefore, comparisons were made

128

AE DC-TR-77-80

to experimental results from other compressors. Trends and

general magnitudes computed by the model were in agreement

with the experimental results.

The response of the compressor to pure radial

distortion was indicated by the model analysis to depend

most strongly on the radial work distribution (radial stage

characteristic variations) built into the compressor.

Response to pure circumferential pressure distortion and

attenuation of the distortion pattern through the compressor

depended on a combination of stage characteristic curve

slope (steepness), and mildly on crossflow distribution.

Response to circumferential distortion patterns with low

pressure region angular extent less than 60 degrees was

indicated to depend strongly on the blading unsteady aero-

dynamic response.

III. RECOMMENDATIONS FOR FURTHER WORK

The modeling technique developed in this work can

be used beneficially in turbine engine compressor design and

development and to assist in solving operational problems

which may arise in field use. There are both improvements

and an extension which can potentially increase its value

further.

Improvements

Stage characteristics. Stage characteristic

information is at the heart of this modeling method.

129

AE DC-TR-77-80

Efforts should be made to obtain improved stage character-

istic information. More attention should be given to

acquiring accurate stage characteristic data during com-

pressor tests. Theoretical methods for computing (or at

least extending the experimental results) should be

developed.

Unsteady cascade aerodynamics. Continued work is

necessary in this area to define the unsteady aerodynamics

of cascades, particularly near and in the stalled operating

regime.

Crossflows. The approximate methods used for

crossflow in the model are somewhat crude. Specific atten-

tion should be given to improving the crossflow represen-

tation. A possible approach is the determination of a

different solution method to solve the governing equations,

including the radial and circumferential momentum equations.

The method would have to be reasonably economic in terms of

computer storage and time requirements. This would also

require determination of a proper representation for the

radial and circumferential forces present in the compressor.

Extension

Post-stall behavior. The model could be extended

to compute the behavior of the compressor after an insta-

bility was entered by incorporation of unsteady aerodynamic

cascade behavior in the deeply stalled regime. Because of

130

AE DC-TTq-77-80

the highly inefficient processes present in the compressor

in deep stall, consideration should also be given to the

metal-to-gas unsteady heat transfer processes. Such an

extension would be valuable for determining methods for

returning the compressor from an unstable to a stable

condition.

131

AE DC-TR-77-80

BIBLIOGRAPHY

i.

.

.

4.

.

.

.

Kimzey, W. F., and E. W. Edmunds. "An Investigation of the Effects of Inlet Flow Density on Compressor Blade Dynamic Stress During Compressor Stall of a YJ93-GE-3 Turbojet Engine in an Altitude Test Cell," Arnold Engineering Development Center TR-66-101, Arnold Axr Force Station, Tennessee, June, 1966.

Montgomery, J. Franklin, III. "The Need for Air Force Engine Stability Margin Testing for Inlet-Engine Interface Definition," Air Force Aero Propulsion Laboratory TR-71-84, Wright-Patterson Air Force Base, Ohio, December, 1971.

Turner, E. E., and M. K. Lake. "Stall Line Determi- nation of a Modified J85-GE-5 Turbojet Compressor," Arnold Engineering Development Center TR-62-150, Arnold Air Force Station, Tennessee, September, 1962.

Calogeras, James E., Charles M. Mehalic, and Paul L. Burstadt. "Experimental Investigation of the Effects of Screen-Induced Total Pressure Dis- tortion on Turbojet Stall Margin," National Aeronautics and Space Administration TMX-2239, Lew~s Research Center, Cleveland, Ohio, March, 1971.

Reynolds, G. G., W. F. Vier, and T. P. Collins. "An Experimental Evaluation of Unsteady Flow Effects on an Axial Compressor," Air Force Aero Propulsion Laboratory TR-73-43, Wright-Patterson Air Force Base, Ohio, July, 1973.

Kimzey, W. F., and R. J. Lewls. "An Experlmental [nvestlgatlon of the Effects of Shock-Induced Turbulent in-Flow on a Turbo3et Engine." Paper presented at the AIAA Second Propulsion Joint Speclalist Conference, Air Force Academy, Colorado Springs, Colorado, June 13-17, 1966.

Plourde, G. A., and B. Brimlow. "Pressure Fluctuatlons Cause Compressor instability," Proceed~nqs of the Air Force Airframe-Propulsion Comparability Symposium. Wright-Patterson Air Force Base, Ohio: Air Force Aero Propulsion Laboratory, 1970. Pp. 576-604.

133

AE 0C-TR-77-80

.

.

i0.

11.

12.

Ward, G. G. "Compressor Stability Assessment Program, Volume I--Dual Spool Compression System Test and Data Analysis," Air Force Aero Propulsion Labora- tory TR-74-I07, Wright-Patterson Air Force Base, Ohio, December, 1974.

Mehalic, Charles M., and Ray A. Lottig. "Steady-State Inlet Temperature Distortion Effects on the Stall Limits of a J85-GE-13 Turbojet Engine," National Aeronautics and Space Administration TMX-2990, Lewis Research Center, Cleveland, Ohio, February, 1974.

Rich, William A., and Robert A. Reale. "Specification Rocket Test for Turbojet/Turbofan Missile Gas Ingestion," Naval Air Propulsion Test Center ATD-197, Naval Air Propulsion Test Center, Trenton, New Jersey, February, 1971.

Boyd, H. E., Jr. "Propulsion System Integration and Test Program (Steady-State), Part VI, YTF-33 Turbofan Engine Tests," Air Force Aero Propulsion Laboratory TR-69-44, Part VI, Wright-Patterson Air Force Base, Ohio, June, 1969.

Gas Turbine Engine Flow Distortion Methodology. Warrendale, Pennsylvania: Society of Automotive Engineers, 1976.

13. Proceedings of the Air Force Airframe-Propulsion Compatibility Symposium. Wright-Patterson Air Force Base, Ohio: Air Force Aero Propulsion Laboratory, 1970.

14. Aircraft Propulsion. Washington, D. C.: National Aeronautics and Space Administration, 1970.

15.

16.

17.

Kimzey, W. F., and S. H. Ellis. "Supersonic Inlet Simulator--A Tool for Simulation of Realistic Engine Entry Flow Conditions." Society of Auto- motive Engineers Paper 740824, Warrendale, Pennsylvania, October, 1974.

Distortion Induced Engine Instability. Neuilly sur Seine, France: Advisory Group for Aerospace Research and Development, 1974.

Whittle, F. "The Turbo-Compressor and the Super- charging of Aero Engines," Royal Aeronautical Society Journal, 35:1047-1074, November, 1931.

134

A E D C -T R-77 -80

18.

19.

20.

21.

22.

23.

24.

25.

26.

27.

28.

Kearton, W. J. Turbo-Blowers and Compressors. London: Sir Isaac Pitman and Sons, Ltd., 1926.

Den Hartog, J. P. Mechanical Vibrations. McGraw-Hill Book Company, Inc., 1940.

New York:

Horlock, J. H. New York: 1973.

Axial Flow Compressors. Huntington, Robert E. Krieger Publishing Company,

Pearson, H., and T. Bowmer. "Surging of Axial Com- pressors," The Aeronautical Quarterly, 1:195-210, November, 1949.

Bullock, Robert 0., Ward W. Wilcox, and Jason J. Moses. "Experimental and Theoretical Studies of Surglng in Continuous-Flow Compressors," National Advisory Committee for Aeronautics Report 861, Aircraft Engine Research Laboratory, Cleveland, Ohio, April, 1946.

Howell, W. T. "Stability of Multi-stage Axial Flow Compressors," The Aeronautical Quarterly, 15:328-356, November, 1964.

Goethert, B. H. "Research and Engineering Studies and Analysis of Fan Engine Stall, Dynamic Interaction with Other Subsystems and Systems Performance," Air Force Aero Propulsion Laboratory TR-70-51, Wright-Patterson Air Force Base, Ohio, July, 1970.

Daniele, C. J., R. J. Blaha, and K. Seldner. "Pre- diction of Axial Flow Instabilities in a Turbojet Engine by Use of a Multistage Compressor Simulation on the Digital Computer," National Aeronautics and Space Administration TMX-3134, Lewis Research Center, Cleveland, Ohio, January, 1975.

Alford, J. S. "Inlet Flow Distortion Index." Paper presented at the International Days of Aeronautical Sciences Meeting, Paris, France, May 27-29, 1957.

Pearson, J. "Wakes in Axial Compressors," Journal of the Royal Aeronautical Society, 63:415-416, July, 1963.

Hercock, R. G., and D. D. Williams. "Aerodynamic Response," Distortion induced Engine Instability. AGARD Lecture Series No. 72. Neuilly sur Seine, France: Advisory Group for Aerospace Research and Development, 1974. Pp. 3-1 - 3-37.

135

AE DC-TR-77-80

29.

30.

31.

32.

33.

34.

35.

36.

Goethert, B. H., and W. F. Kimzey. "Effect of High Frequency Fluctuations of Inlet Flow on Compressor Stall." Paper presented at the 32nd Meeting of the Propulsion and Energetics Panel of AGARD, Toulouse, France, September 9-13, 1968.

Reid, C. "The Response of Axial Flow Compressors to Intake Flow Distortion." American Society of Mechanical Engineers Paper No. 69-GT-29, presented at the Gas Turbine Conference and Products Show, Cleveland, Ohio, March, 1969.

Calogeras, James E., Charles M. Mehalic, and Paul L. Burstadt. "Experimental Investigation of the Effect of Screen-Induced Total Pressure Dis- tortion on Turbojet Stall Margin," National Aeronautics and Space Administration TMX-2239, Lewis Research Center, Cleveland, Ohio, March, 1971.

Goethert, B. H., and K. C. Reddy. "Unsteady Aero- dynamics of Rotor Blades of a Compressor Under Distorted Flow Conditions." Paper presented at AGARD Fluid Dynamics Panel Specialist Meeting on Aerodynamic Interference, Silver Springs, Maryland, September 28-30, 1970.

Peacock, R.E. "Unsteady Response of Compressor Rotors at a 90 ° Circumferential Pressure Distortion," Proceedings of a Workshop on Unstead~ Flows in Jet Engines Held at United Aircraft Research Labora- tories, F. O. Carta, editor. West Lafayette, Indiana: Purdue University, Project Squid Head- quarters, 1974. Pp. 193-210.

Korn, J. A. "Estimated Effect of Circumferential Dis- tortlon on Axial Compressors Using Parallel Compressor Theory and Dynamic Stall Delay." American Institute of Aeronautics and Astronautics Paper No. 74-33, presented at 1974 AIAA Aerospace Sciences Meeting, Washington, D. C., January- February, 1974.

Carta, F. O. "An Experimental Study on the Aerodynamic Response of a Subsonic Cascade Oscillating Near Stall," Project Squid Headquarters TR-UTRC-2-PU, West Lafayette, Indiana, July, 1976.

Melick, H. C. "Analysis of Inlet Flow Distortion and Turbulence Effects on Compressor Stability," National Aeronautics and Space Ad~,inlstration CR-ii4577, Moffett Field, California, March, 1973.

136

AE DC-TR-7?-80

37.

38.

39.

40.

41.

42.

43.

44.

Gabriel, David S., Lewis E. Wailner, and Robert J. Lubick. "Some Effects of Transients in Inlet Pressure and Temperature on Turbojet Engines." Paper presented at the 25th Annual Meeting of the Institute of the Aeronautical Sciences, New York, January 28-31, 1957.

Kimzey, William F. "The Effects of Unsteady, Non- uniform Flow on Axial Flow Compressor Stall Characteristics." Unpublished Master's thesis, The University of Tennessee, Knoxville, 1966.

Kuhlberg, J. F., D. E. Sheppard, and E. O. King. "The Dynamic Simulation of Turbine Engine Com- pressors." American Institute of Aeronautics and Astronautics Paper No. 69-486, presented at the AIAA/SAE Joint Propulsion Specialist Conference, Cleveland, Ohio, June, 1969.

Willoh, Ross G., and Kurt Seldner. "Multistage Compressor Simulation Applied to the Prediction of Axial Flow Instabilities," National Aeronautics and Space Administration TMX-1880, Lewis Research Center, Cleveland, Ohio, September, 1969.

Jacocks, J. L. "Prediction of Maximum Time Variant Inlet Distortion Levels," Proceedings of a Workshop on Unsteady Flow in Jet Engines Held at United Aircraft Research Laboratories, F. O. Carta, editor. West Lafayette, Indiana: Purdue University, Project Squid Headquarters, 1974. Pp. 377-390.

Kimzey, William F., and Milton W. McIlveen. "Analysis and Synthesis of Distorted and Unsteady Turboengine Inlet Flow Fields." American Institute of Aero- nautics and Astronautics Paper No. 71-668, presented at the AIAA Seventh Propulsion Joint Specialist Conference, Salt Lake City, Utah, June 14-18, 1971.

Korn, James A. "Propulsion System Flow Stability Program (Dynamic) Part XI, Lift Engine Compressor Tests and Response to Steady-State Distortion," Air Force Aero Propulsion Laboratory TR-68-142, Part XI, Air Force Systems Command, Wright- Patterson Air Force Base, Ohio, December, 1968.

Voit, Charles H. "Investigation of a High-Pressure Ratio Eight-Stage Axial-Flow Research Compressor with Two Transonic Inlet Stages, I--Aerodynamic Design," National Advisory Committee for Aero- nautics Report RM-E53124, Lewis Flight Propulsion Laboratory, Cleveland, Ohio, December, 1953.

137

AE DC-TR-77-80

45.

46.

47.

48.

49.

50.

51.

52.

53.

Geye, Richard P., Ray E. Budinger, and Charles H. Volt. "Investigation of High-Pressure-Ratio Eight-Stage Axial-Flow Research Compressor with Two Transonic Inlet Stages, II--Preliminary Analysis of Overall Performance," National Advisory Committee for Aero- nautics Report RM-E53J06, Lewis Flight Propulsion Laboratory, Cleveland, Ohio, December, 1953.

Voit, Charles H., and Richard P. Geye. "Investigation of a High-Pressure-Ratio Eight-Stage Axial-Flow Research Compressor with Two Transonic Inlet Stages, III--Individual Stage Performance Charac- teristics," National Advisory Committee for Aeronautics Report RM-E54HI7, Lewis Flight Pro- pulsion Laboratory, Cleveland, Ohio, November, 1954.

Milner, Edward J., and Leon M. Wenzel. "Performance of a J85-13 Compressor with Clean and Distorted Inlet Flow," National Aeronautics and Space Administra- tion TMX-3304, Lewis Research Center, Cleveland, Ohio, December, 1975.

Bruce, E. P., and R. E. Henderson. "Axial Flow Rotor Unsteady Response to Circumferential Inflow Dis- tortions," Project Squid Headquarters TR-PSU-13-PU, West Lafayette, Indiana, September, 1975.

Ostdiek, Francis Richard. "A Cascade in Unsteady Flow." Ph.D. dissertation, Ohio State University, Columbus, 1975.

Carnahan, Brice, H. A. Luther, and James O. Wilkes. Applied Numerical Methods. New York: John Wiley and Sons, Inc., 1969.

Roache, Patrick J. Computational Fluid D~namics. Albuquerque, New Mexico: Hermosa Publishers, 1972.

Wallhagen, Robert E., Francis J. Paulovich, and Lucille C. Geyser. "Dynamics of a Supersonic Inlet Engine Combination Subjected to Disturbances in Fuel Flow and Inlet Overboard Bypass Airflow," National Aeronautics and Space Administration TMX-2558, Lewis Research Center, Cleveland, Ohio, May, 1972.

Milner, Edward J., Leon M. Wenzel, and Francis J. Paulovich. "Frequency Response of an Axial-Flow Compressor Exposed to Inlet Pressure Perturbations," National Aeronautics and Space Administration TMX-3012, Lewis Research Center, Cleveland, Ohio, April, 1974.

138

AE DC-TR-77-80

54.

55.

56.

Plourde, G. A., and A. If. Stenning. "The Attenuation of Circumferential Inlet Distortion in Multistage Compressors." American Institute of Aeronautics and Astronautics Paper No. 67-415, presented at the AIAA Third Joint Propulsion Specialist Conference, Washington, D. C., June, 1967.

Aerodynamic Design of Axial-Flow Compressors. Revised. Washington, D. C.: National Aeronautics and Space Administration, 1965.

Langston, C. E. "Distortion Tolerance - By Design Instead of By Accident." Paper presented at the International Gas Turbine Conference and Products Show, Cleveland, Ohio, March 10-13, 1969.

139

AEDC-TR-77-80

APPENDIX A

FIGUR[:S FOR CHAPTERS I THROUGH VI

14|

AEDC-TR-77-80

Compressor Turbine

/ - - - A f t e r b u r n e r

/~=- Jet N o z z l e

a. Single-spool afterburning turbojet

- ~ompressor

o

b. Dual-spool augmented (afterburning) turbofan, bypass ratio ~ 1:1

Fan

c.

Figure 1.

~ Rear Compressor

Dual-spool high bypass ratio turbofan, bypass ratio ~ 6:1

Typical aircraft gas turbine compression system configurations.

142

4~

8 . 0

0 .4 &J

¢ 6 . 0 S.,

r.4

0

o 4 0 .

E 0

2,0

25

J

Aerodynamic ~ 1 0 0 ~ % Stability Limit Line, / "Stall Line", or

. .

~--Operating N/~ = 80% Line

(Corrected Rotor Speed, percent of Design)

i

3 0 35 4 0 45

Corrected Airflow, W/e 6

Figure 2. Typical compressor map. m

.o -4

& o

AE DC-TR-77.80

9 Flight .o Bach A l t i t u d e ,

Number f t ~ , ~

~ 8 o%7W-~ ~o--7~7~-- ~ A

== 1°s 3o,~oo = o,oo

8

3 1 80% I I , I , I , I 28 30 34 38 42 46

C o r r e c t e d Airflow ~W'v l b m / s e c ' 6

a. Compressor map operating regions for a single-spool turbojet engine operating at constant mechanical speed

7 x 103-

60-- L

6 4 o - < ~0~,

"~ 3 0

0 ~ i , , / I , /, , , I 0 0 .4 0 . 8 1 .2 1 .6 2 .0

Flight Mach Number

b. Variation of engine inlet total temperature with flight condition

Figure 3. Effect of flight altitude and Bach number on compressor operating region.

144

AEOC-TR-77-80

0

a

o

E o

8

6

5 35

REI

- 1 . 0

t-108%

0%

5%

\ 90% m i

i I i I I i i i i I 40 45

W/8 Ibm/sec Corrected Airflow, -~--,

c. Influence of Reynolds number on compressor stability limit [3]

70 x 103 r R E y

5 0

"~ 3 0

10

o ' I I i I / I ~ [

0 0 . 4 0 . 8 z . 2 z . 6 2 . 0

F l i g h t M a c h N u m b e r

d. Variation of Reynolds number with flight condition

Figure 3. (Continued).

145

AEDC-TR-77-80

Dis to r t i on Screen Pa t te rn

8 -- ~ H i g h Pressure / ~ P H - PL -

o I//f ~ eeslon, P. / _~=~- -- = 12.5~ FX C~;'/A~Lo. / _s--, , ' ,

" T~'/A ~ " ~ / / / / / / / A Pressure / _~" s " " I at N//~ = 100~

6 -- / \ /~ "---Stability Li,-it / / ~ ~ with Dis to r t i on

~ . . . . / / / -

I ~ ~-- Undis tor ted / S t a b i l i t y Limit

87% 4

32 34 36 38 40 42 44 46

w~ Corrected Airflow, T ' lbm/sec

a. Influence of steady, screen generated distortion on turbojet compressor stability limit [4]

O ~4

m

k

W m

k

2.8

2.4

2.0

1.6

1.2 I00

|I _L ,~ o.., p,v,, / A

/ / ~ ~-- S t a b i l i t y Limit / / ~ ~ wlth AP = 0

f s ~" ~ ~ - S t a b i l l t y Limit with I ~ AP = 0.5 Pavg

N I / ~ = s0~ , , , , l I l l l I I I I I

200 300

Corrected Airflow, W#, lbm/sec

b. Influence of uniform time varying flow at the inlet of a turbofan engine [5]

Figure 4. Typical influences of external disturbances on compressor stability.

146

AEDC-TR-77-80

~I _L s , - ~ ~k-A--£~-~i-~-~p . 0.1s Pav

¢~ g o L = ~o~ avg

,, /.,,,, 3J-Stability Limit withJ ~ l

I Combined Uns teady | _ 80% 2 r and Distorted Flow 75%

,3 l / I I I I I I i00 120 140 160 180 200 220

C o r r e c t e d A i r f l o w , W ~ , l b m / s e c

I 240

c. Influence of time-variant distortion on turbojet engine compressor stability limit [6]

o 8 ~a

= 7

k

= 6

5

o 4 W

k 3 E o 2

d.

1 60

- / ~ - \ .... ~ io.ooo

_ ~ ~ ~ - - S t a b l l i t y Limit wi th rpm S t e a d y , Uni fo rm I n f l o w

7 ,000 S t a b l l i t y Limit wl th Combined rpm Unsteady and Distorted Flow

l I l I I I f 70 80 90 100 110 120 130

C o r r e c t e d A i r f l o w . W_~, l b m / s e c

Influence of time-variant distortion on the low-pressure compressor of a turbofan engine [7, 8]

Figure 4. (Continued).

i

147

AE DC-TR-77-80

3 . 6 vStability Limit w i t h \ S t e a d y Uni form I n f l o w

3.2 7 % / " \ \ w i t h Distort*on V ~ ' ~ \ \ and U n s t e a d y / .., '- \\ ~% I n f l o w from

~ 3 . 0 ~ " ./" \\ ~ Discharge of "~, ./*" \\ ~. U p s t r e a m

~ 2.8 ~, ~ I\; OmpressOr

~ 2.e ~\ ~ N/.~ = zo, soo rpm

o 2 . 4 %\% 10,000 rpm \

9 , 5 0 0 rpm 2 . 2 I I I I I

20 22 24 26 28 30

C o r r e c t e d A i r f l o w , W_~, l b m / s e c

e. Influence of unsteady and nonuniform flow from the low-pressure compressor on high-pressure compressor stability limit in a turbofan engine [8]

O

ee

L

~q

O

O

¢ L

E O

9

8

7

6

5

4 35

- ~ / L o w T e m p e r a t u r e Re g ion , T L / / / / / ~ (T H - TL) /Tavg = 19%

~ / ~ j r " H i g h T e m p e r a t u r e R e g z o n , T H _ ~ / ' ~ . . . _ g + o ~ ,

-- ~ " - - ~ t a b l l l t y L i m i t / ~ ~ wlth Unzform

t.. - ~ ' ~ - - ~ ' ' ~ . ~ " N / / '~ = 96.2% ~ . N/CO = 93.1%

- S t a b x l x t y Lzmi t w i t h Nonunzform T e m p e r a t u r e

K , z J I i , , , I 40 45

C o r r e c t e d A i r f l o w , ~ , l b m / s e c

f. Influence of nonuniform total temperature on a turbojet engine compressor [9]

Figure 4. (Continued).

148

A E D C-T R-77 -80

C o m p r e s s o r - ~

~-~ \ ~

¢ 0 P2

s Throttle

~ - - - - ~ T h r o t t l e Closed

l l

Throt t le Open

Cq n.

d +J

o ~=

T..

Z'

/,

C

D

C

B

Loss Character- istics of Throttle

u o m p r e s s o r Character- i s t i c

Flow Rate

Figure 5. Surge analysis based on compressor and throttle curve shapes.

149

AE DC-TR-77-80

F 8tarot Blade

I I

F !

--'Y'------"-- --~lnletL Guide Vane -- --

U

eee-r'~- - ...... " / " ....... "~ ii ~ G VT ~ °° ee ° ' ' : el o o o I m l l b ° ° e ~ ° ° el'

D2ntance Into Compressor

Figure 6. Representative axial flow compressor arrangement and operation.

150

A E D C - T R - 7 7 - 8 0

Turbine Guide Vanes 7

Inlet Duct Compressor Burner /

a. Physical system modeled

~ p E f f e c t i v e Axial Force . ~ . B l e e d Flow Per er Unit Length, F(z , t ) UnitwB(z,t)Length

t P S ( t )

, _ . ,

#S ha f t Work and Heat Per Unit r ~ Length, WS(z,t) and Q(z,t)

z

b. Overall system control volume

_ _ ~ dz ~ ~ ~ _ _ _ _ _ BA A A + ~-~ dz

w __I=!F I V ~w PS dz _ L :- W + ~-~ dz PS ÷ ~--P~. dz

U =- U* ~--~ dz H : BH

H + ~-~ dz

W8 dz + Qdz

c. Elemental control volume

Figure 7. Physical system modeled and control volumes.

151

AE DC-TR-77-80

Ve 2

F e

Y 1

S F l u i d E lement

a. Rotating fluid element

hord L ine F8

V

U

mr = Uwh

V 8 = Uwh + Vsw

V sw

(a + ~)

Axlal Directlon

F 8 = L cos (a + A) + D sxn (a + A)

b. Blade element velocities and forces

Figure 8. Rotor fluid element velocities and forces.

152

AEDC-TR-77-80

_~ :~,

, -

¢/2

~o

4~

~ c

o3

1.2

I 0

0.8

0.6

0.4

0.2

1 28

1 24

1 20

1 16

1 12

o~/

i 0

_ J

- - - - C

~ t

, !

" i Io

1 08

1.5 [ i

1.3 --. j ]

~.~ ~ 'o l I 0 . 2 0 . 3 0.4

_i [

I !

I

!

; I ,

I I i

~ N U

! l ' I

0.5 0.6 0.7

Flow Coefficient, ¢

Sv.~

O rl

O

V

m ~ m

0.8

Corrected Speed, Percent DesiEn

30 50 60 70 80 90

100 D e s z g n

Figure 9. Representative stage characteristics [46].

153

4a

Steady Flow

~P

Unsteady Flow

S tage C h a r a c t e r i s t i c Curves

~T

~P ss /

s Sp • l

l

~T, ~P

I I

l S

I I l 1 l 1 l l

E q u i v a l e n t L i f t C o e f f i c i e n t Curves

c~ c~ ~ c~ / \

/ /

Figure i0. Steady-state and unsteady stage characteristic behavior.

> m 0 ? "4

& O

Stagger k = 50°I~. i ~ ~ _ ~ Cascade a/c : ~°~Io~

u o. S I - - ~ k \ X ¢ - 1.o

-- 0 ii 1 3~ k -- -_~._C Ic~l~x Vrel [Cllss 0

0"0 , - _ .

a ~ 0 2 4 k 6 8 I0

Oscillating Stream, Stationary Folls

k a[ r]

U Actua~ I00- l

T Y/? c 'V / ~J. '_c_" s. ..,........ r," .7- ~ , - . ~ ' L-quasi-Steady Stat(,

O. O. 02 O. 2 2 20

k Oscillating Foils.

Steady Stream

Figure 11. Unsteady airfoil and cascade aerodynamics.

m 0 9 -4

m 0

t , / t o~

I C ~ l m a x -r~-[-ss

1 . 2

1 . 0

0 . 8

0 . 6

0 . 4

0 . 2

0

0 . 0 2

On cP

[]

= 50 ° a / c = 1 , 3 - - X O

% B r u c e a n d H e n d e r s o n . 14B]

0 ~ = 4 s o a / ~ = 1 . ~ s 3 \ D X = 55 ° a / C = 1 . 3 5 3

=__c_c N o t e B r u c e and H e n d e r s o n D a t a k = V r e l N o r m a l i z e d t o A v e r a g e o f Co V a l u e s a t L o w e s t V a l u e o f k (% 0 .6 ~

C ffi F u l l B l a d e C h o r d L e n g t h

I I i I I I [ I I I I J i i i j I I I 0.04 0.06 0.1 0.2 0.4 0.6 1.0 2.0 4.0

k

I I I I I 6.0 10.0

Figure 12. Comparison of theoretical and experimental unsteady cascade lift coefficients.

m o n "4 3

& 0

t ~

1 . 2

~00

OH 1.1

oe o

• ,4 d ~ 1.0 o o

• , 4 4-~

t ~ e l

~ 0 . 9 O ~

O ~ • , 4 . - I

G d ~

0 . 8

O s t d 2 e k D a t a f o r 16 < Z < 24 d e E , a / c = 0 . 8 1 8 5

B r u c e and H e n d e r s o n D a t a f o r = 45 d e g , a / c = 1 . 3 5 3

<>

O

O s t d i e k C a s c a d e D a t a , [49 ]

0 0 . 0 0 2 ( k )

D 0.044

[~ ~ O. 080 I',, ~ o .12o

C3 o,1 o <>

B r u c e a n d ( ~ 0 H e n d e r s o n , [ 48 ]

E]

0 . 5 6 ( k )

1 . 1

2 . 4

2.9

0

I I

2

Figure 13.

I l . I I I I 4 6 8

Mean A n g l e o f A t t a c k , ~m' d e g

10 12

Variation of dynamic lift coefficient ratio with mean angle of attack.

m

o 0

~s

m 0

1 . 0

G o e t h e r t - R e d d y S o l u t i o n , [ 32 ]

fll

~3 ?

& 0

0 . 8 O F i r s t O r d e r A p p r o x i m a t i o n

IC£[ma x

IC~Iss

0 . 6

0 . 4

0 . 2

0

Ic~{max = . f IC~lss ~ 1 + ( e k ) 2

e = 0 . 3 1 0 6

A p p r o x i m a t e a a / c = 1 . 0 R e s p o n s e E q u a t i o n

A ffi 50 ° C£ C£ss ~ + -- =

T T

eC T = Vrel

0 0

k = _ _

I I I I 2 4 6 8

R e d u c e d F r e q u e n c y , k

10

2 . f C

V r e l

Figure 14. Unsteady lift coefficient magnitude ratio.

AEDC-TR-77-80

Plane Location Designations

im = i - 1 Stator--~

t m

q

C/_.o.o. /

ip= i +I

-A

a. Physical system

ira=i- 1 ip= i + 1

Wim

PSi m

Him

v

v

W i PS i H i

Axial i • Z • v im z. z.

Locations m l p

Wip

P S i p

Hip

b. Computational planes

Figure 15. Computational plane designations for finite difference solutions•

159

0

Cumpute V a r i a b l e s at Each P lane

• , Entry Plane C o n d i t i o n s

U I " M(PI,TI.WI.A I ) Eq. (75) TS 1 • TS(TI.U 1) Eq. ( 7 6 )

PSI = PS(P1,M 1) Eq. ( 7 7 )

o I = o(PSI ,T81) Eq. (78) XI i X(PSI.MI ) Eq. ( 7 9 )

I n t e r i o r Plane Values

UI i U(Wl,01AI) Eq. (80) TS i " T S ( X i , 0 | , U i ) Eq. (81) P5 i = PS(oi ,TS I ) Eq. (82)

T i " T(TSA,U I ) Eq, (83)

Pl = P i ( T i ' T S I I Eq. (84) M i " M(UI,TS i ) Eq. (86)

IMP i - IMP(PSI.Ai.M i ) Eq. (87)

8 i = B ( g i , T i . c p) Eq. (88) m m m m m m m ~ N m m N m m m m m r a m ,

E x i t P l ane C o n d i t i o n s

I f P i n / P e x i t ~ ( P / P S l c r l t Eq. (89) Min .- 1 .0

I f P l n / P e x i t < ( P / P S ) c r l t Eq. (90) PSin ~ P e x i t

Min = Min(X.Pex i t ) Eq. (81) TSin " TS(PSin. Oin) Eq. (921

Uin = U(Min,TSln) 8q. (93)

g in " g i n ( O i n ' U i n ' A e x l t ) Eq. (94)

At t - 0

At t ~ 0

I n i t i a l C o n d i t i o n s |

l ( o A 1 i ( t i 01, W i ( ; ' - 01, (XA1i( t = 01, * i ( t - 01 " lnpu

Runge-Kuttn I n t e g r a t i o n

I

( P A ) i ( t + At) = 8 ( a ( p A ) l l | t , ( p A l l ( t ) )

| i ( t ÷ At) - 8 ( a V i l a t . g l ( t l ) (XA)l(t + A t ) - 8(~ (XA) l /~ t , (XA11(t11

$ i ( t + At) - f ( ~ $ i / ~ t , O l ( t ) )

T Compute Time D e r i v a t i v e s

] a ( o A ) i / ~ t - f l ( g i m , W i , Wip.WB i )

~wi /~ t = f2(IMPim, IMPi. IMPip 0F i ) ~(XA) I /a t - f 3 ( H i m , H f . H i p . g S l l

~ e i / a t = t 4 ( e i . $ n n l )

T Compute Forces and gork

Compressor

# i " # i ( U i ' Va i l e i - e l ( I l l ~ q l n i ( t i ) } Curve F i t s

PR i = P R ( e i . N i . T l ) TR i - T R ( P 8 1 . , i )

Y 1 = F i ( P R i . l i , A i p ) 88 i - gSi(TRA,HI)

Duc___~t

F i = Fi(PSi, M i, A i )

Eq. "(63) ,Eq. (104) Eq. ( 6 4 ) . Eq. (105) Sq. ( 6 5 ) . Kq. (1081 Eq. (581

• qo ( e e )

Eq. (67) Eq. (68) Eq. (701 8q. ( 8 9 )

Eq. (111)

I n p u t s (Boundary C o n d i t i o n s )

P l i Inpu t P e x i t = Inpu t

T 1 " Inpu t A e x i t = I n p u t Qi i Inpu t N l o Inpu t

m o ?

& 0

Figure 16. Solution procedure for one-dimensional, time-dependent model.

AE DC-T R-77-80

1 2 3

® ®

J

4 5

® ®

Compressor

a. Allison XC-I compressor

1 2 3 4 5 6 7 8 9 I0 11 12 13 14 15 16

@

Duct I_

Turbine Nozzle

b. NACA-8 and J85-13 compressors

Figure 17. Station locations (computation planes) for one-dimensional, time-dependent models.

16l

AE DC-T R-77-80

4 . 0

~ 3 . 0 -

GI

2 . 0 0

4.0 m

k

M

k ~

= 3.5--

o - o

L t

o o 3 . 0

0

o 35 -

~ m

u s 34 --

k o -

33 --

E O

32

0

P l = 1 4 . 7 p s L a

T I = 5 1 8 . 7 OR

t I I i 10 20

I I I 10 20

I I I I I 10 20

P l a n d T 1

H e l d C o n s t a n t

I I I I I I I I 30 4 0 50 60

TLme, msec

/ F l o w B r e a k d o w n ~ " I n L t t a t L o n ~ i

I

I I I ] I I I' I 30 40 50 60

Time, msec [

[

!

I I i I I I ~ I 30 40 50

TLme, m s e c

Figure 18. XC-I compressor model loaded to stall,

N//~= 9 0 % .

6 0

162

AEDC-T R-77-80

o,

j ~.oI-- t . . . . . . . . . . - ~ \ ~

52 53 54 55 56 57 Time. msec 1.1-

1.0

0 .9

s ~

- O Stage 1 Q S tage 2 0 Stage 3

-- ~ S tage 4

0 .8 I I 52 53

4.0 m

Stage 2 ~ 1 t _ _ ~ / O ~ /

I I I I l I I 54 55 56 57 Time. msec

3.0

2.0

0

o

1.o i I I 52 53

Stage 3 I

Stage 4 Entry Entry~/

(S t age 3 E x i t ) - - ~ 0

O- 0-0-.0_ O_ 0.-0-0-- 0/~ Stage 2 E n t r y - - ~ ~ .~._ D___~__O.__O.__DImD..O.__D~D O-~

I I I I I I ~ ' 54 55 56 57 Time, msec

Figure 19. Flow breakdown process leading to instability, XC-1 compressor, N//~ = 90%.

163

AE DC-T R-77-80

4 . 0

1-4

J qq 4~

3 . 5

¢1

1,4

$4 0 W

o 3 .0

2.5

i ~ r p e r c e n t lmental , ~ Stability Limit (from [43])

- - - - - - . 4 m

-- I 0.6 Percent

~----Flow Breakdown from Model

Model Computed 90-Percent Speed Character- istic

Flow Breakdown from Experimental Data

sym

0 Experimental Data, N/~/e = 90 Percent (from [43])

C

31

I I I I 1 I I I I J 32 33 34 35 36

C o r r e c t e d A i r f l o w , Wv~/6, l bm/sec

Figure 2 0 . Comparison of model computed 90% speed charac- t e r i s t i c and stability limit with experimental measurements, XC-I compressor.

164

0.50

tn

0.45

0 . 4 0

0.35 t ~

=~ 0 . 3 0

,==

• -~ 0 . 2 5 N N O

Z

® 0 . 2 0 ,4

JD k ,

[ - 0 . 1 5

O . 1 0

0.05

--- . . . .

P l = 10 p s l a

T 1 - 5 1 8 . 7 ° R

0 . 1 0 . 2 0 . 3 0 . 4 0 . 5 0 . 6 T l m , sec

Figure 21.

a. Imposed nozzle area reduction.

NACA-8 compressor model loaded to stall, N/~ = 90%.

m o O

& O

1 2

1 1

1 0

dl n.

"A 9j e..4

n .

1'4

g

~ 7 k I

GO III fO J4 ~ _ _

A. 6

h 0 B m 0 '-- I

[] 0

U 4 '

3

2

J

J i

f

I J

I I

0 . 1 0 . 2 i i

0 . 3 0 . 4 0 . 5 0 . 6

T i m e , Bec

b . Compressor pressure ratio

Figure 21. (Continued}.

m O

& O

1.20

O~ -,3

,.4 4~

~1 ~

o ,...4

4J

Q;

k 'D

r j

k o

w

o,

o (.)

] , 1 6

1 . 1 2

1 . 0 8

I , 0 4

1 . 0 0

0 . 9 6

0 . 9 2

O , 8 8

\

0 . 8 4

I II H |

0.80

0 0.I 0 . 2 0 . 3

Time, sec

c. Corrected airflow.

Figure 21. (Continued}.

0.4 0.5 0.6

m O

~s

00 o

A E D C-T R-77-80

9 . 0 m

8.0 --

I

7 . 0 -

n .

B

a 6 . 0 -

L

m Ui 0

~' 5 . 0 - h 13

10 - - qo L. eL

E 0 4 . 0 -

3 . 0 -

2 .0 2O

Figure 22.

E x p e r i m e n t a l S t a b i l x t y Limit ( f r o m [45] )

2.2% - - - ~ "

,1 L.~ ~,

#

T'-'II'-/

t'-~/~°~ °

L o 0

0

0 0

OoN/~ = 90%

°o 0 P o i n t s f rom E x p e r i -

m e n t a l Compres so r map ( f r o m [45] )

Computed C h a r a c t e r i s t i c s and S t a b i l i t y Limits f rom Model

11 °

I I I I I I I 30 40 50 60

w ~ C o r r e c t e d A i r f l o w , T ' I b m / s e c

I j 70

Comparison of model computed speed character- istics and stability limits with experimental measurements, NACA-8 compressor.

168

AE DC-T R-77-80

0 . 8 [ -

| - - - 0 - - E x p e r i m e n t a l , [45] ~ . . . - - - - - - 0 L - - 1 3 - - " Computed _ _ ~ ~ o. s, , L - - - - -

0 .4 L - x- s t - - a l l - ~ - ' ~ ~ s t a g e 1 p_-- '~ o . 2 ~ I i I I I • ~ 50 60 70 80 90 100

i 0 8 F N/ /~ . p e r c e n t

/ ~ ~ 0

- - 0 4 . - - 0 " S t a g e 2

= 0 2 l I I I I I s 50 60 70 s o s o z o o 0

~J N/4e, p e r c e n t -~ 0 8 -- ed

4J

~ 0 6

~ 0 4 0 0

0 . 2 c 50

~ 0 . 8 -

D

~ " " " S t a g e 3

I I I I I 60 70 80 90 100

N/V~, p e r c e n t

0.S -- _--n-------~---~~------~ ~ - - - - - - u -- - - - - - ~ - -

0.4 S tag e 4

0 . 2 I I 1 i I 50 60 70 80 90 100

N/V~, p e r c e n t

a. Stages 1 through 4

Figure 23. Comparison of model computed flow coefficients at compressor stability limit with experimental measurements, NACA-8 compressor.

169

A EDC-TR-77-80

0.8 - s_~

- - - - 0 - - - - E x p e r i m e n t a l , [ 4 5 ] 0 . 6 - - - u - - Computed

.~o~..~. ...... ~. ...... .-° ~-~ 0 . 4 S t a g e 5

o . ~ I I I I I 50 60 70 80 90 100

N/a, p e r c e n t 0 . 8 ~

0 . 6 S t a g e 6

• r ~--'~=~---'~ .......

0 " 4 1 ~ " ' - - ¢ s t a l l " " - - - - L e s s Than 0 . 4

0 . 2 / [ I I ] I 50 60 70 80 90 I00

= N / a , p e r c e n t 0.8- - O ~4 qq

0.6

~o 0.4 -...4)

0.2 [ a 50 100 4a

G3

-- , S t a g e 7

~ - " - u - - = ~ - - - - ~ . . R .

- --'~¢stall

I I I I 60 70 80 90

N / / e , p e r c e n t O.S -

0 . 6

~ ¢ s t a l l I

60

S t a g e 8

-- w ~ m m ~ R D m m i m 0 0 . 4

L e s s Than 0 . 4

0 . 2 I I I I 50 70 80 90 100

N/V~, p e r c e n t

b. Stages 5 through 8

Figure 23. (Continued).

170

AEDC-TR-77-80

5 . 8

5 . 4

5 . 0 -

4 . 6 -

¢ L

if)

4.2 - - C .

L

D

L

~- 3 . 8 -

3 . 4 -

3 . 0

2 6

/ S~an111ty Ll.mz t / F ~

~ / Characterlstic

N / d ' ~ = 8 7 ~

O Points from Experimental Compressor Map (from ,47])

Computed Characterxstic and Stabillty L]mlt

~ / ~ = 8o~

i I I I I I I I l 28 30 32 34 36

~'~ Corrected Azrflow. - - - ' q - . . Ibm/sec

Figure 24. Comparison of model computed speed character- istics and stability limits with experimental measurements, J85-13 compressor.

171

1 8 . 0 0

1 6 . 6 0 ] l,n IA n lll w

15.20 A ' I

i ~ 10 , in/I ~i II II ~,,..,o ~ ~ II I 1 !1 II II., - ' " . .

~ ,,.oo , , A f I I l . . l l l l l l l . . "" ~, \i I l l Ill II lal " 9 . 6 0 . .

, v ~ I l l i1111 IIII 8 . 2 0 • • , ,

u I/I lit II Illl "~ 6 . 8 0 • ,, , , _

~ uilJ i/ii - , f : 2 0 H z O

U 5 . 4 0 . .

.oo 1 v l l 0 . 1 5 0 . 2 5 0 . 3 5 0 . 4 5 0 . 5 5 0 . 6 5 0 . 7 5 0 . 8 5

TAme, s e ¢

a. Inlet total pressure

m

0 ?

~ J

oo o

Figure 25. Oscillating inflow loading to stall, NACA-8 compressor, ~I/~= 7 o ~ .

%

J.J

o

6,80

6 . 4 0

i~ O0

rj. 60

5.20

4 80

4 . 4 0

4.00

2.80s 0

i~N//O = 9o% I , - - -Dece] t o

70% Speed

!

f = 20 Hz

-l ;

~" 70%ipeed

O. 0.2

/I/

0 3 0 . 4 0 . 5 0 6

Time, $ec

0.7 0 8 09

b. Compressor pressure ratio

Figure 25. (Continued).

> m o

~4 & 0

--.I

1 O0

O. 90

0.80

~o7o

0 GO

o . 50

.7

k .~ O. ,10

t. 0 3 0

k

E 0 O . 2 0

0 . 1 0

0 0 1 5

nP *I

Ill II Ill II II II/

f - 2 0 I I z

ItlllIJ

V, 0.25 0.35 0.45 0.55 0.65 0.75 0.85

Time, sec

c. Compressor entry airflow

> m o n

m o

Figure 25. (Continued).

..j ta-,

0.6 01 I

0.5 1 0 . 4

~ 0 . 3 / S t a b i l i t y L i m i t

l ~, o . s

: .- 0 . 4 - 0 2 a t S t e a d y S t a

S t a b i l i t y L i m i t

0 . 3 - f =

0 . 2 I I I , I I ., I I O. 15 0 . 2 5 0 . 3 5 0 . 4 5 0 . 5 5 0 . 6 5 0 . 7 5 0 . 8 5

Time, se¢

d. Flow coefficients, stages 1 and 2

Figure 25. (Continued).

m O

~J & O

.,.j

0 .6 %3

0.5 .P

O.,l

~ 0.3

o (J

0.6 ~ %4 / " - ~ 4 at Steady-State - ' ~

0.5 u)

0.4

0.3

0 . 2 1 I i I I , I I ~1 O. 15 0 . 2 5 0 . 3 5 0 .45 0 .55 0 . 6 5 0 . 7 5 0 . 8 5

Time, Rec

e. Flow coefficients, stages 3 and 4

> m O ? 3

m O

Figure 25. (Continued).

~4 ..j

0.6[-- ~ ' 5 at S teady-S ta te , I / Stabl I tty Limit 5 - - ' ~ _

0.5

0.4

0.6 / " - -*6 at S t eady-S ta t e 6~_

e6s t~ l l Less T:an 0 . 3 I I O. 15 0.25 O. 35 0.45 0.55 0.65 0.75 0.85

Time, sec

f. Flow coefficients, stages 5 and 6

Figure 25. (Continued). > m 0 n

$

oo

u

o

0 . 7

0 , 6

0 . 5

0 . 4

~p --@7 a t S t e a d y - S t a t e S t a b i l i t y L s m i t

. 7 s t a l I

0S 0 . 7 - - ,_. . . . . . .0 s a t S t e a d y - S t a t e ~ .

~o / S t a b i l i t y L i m i t

0 . 6

0 . 5

0 . 4 d 0 . 1 5 0 . 2 S 0 . 3 5 0 . 4 S 0 . S S 0 65 0 . 7 5 O .S5

T ime . s e c

m O n

m o

g. Flow coefficients, stages 7 and 8

Figure 25. (Continued).

~D

+J

O

O U

4~

~0

¢d

2.0

1.5

1.0 0.9 0.8 0.7 0.6 0.5

0.4

0.3

0.2

0.1

A 1-7-

/Rotor Blade /Roduced ~Frequency ~ time

-k I = 0 . 0 6

- ~ - O - C a l c u l a t e d w i t h Q u a s i - Steady-State Assumption

0.11~ -Q-Calculated with Goethert- Reddy Unsteady Cascade Correction

I\ $ Compressor Unstable I ~ / i n this Region (Stalls)

Stable Compressor ~ NI 4~ = 70% O p e r a t i o n i n t h i s ~ Reglon (No Stalls) ~b

I,, k I = 0.88 v

I , | , I , I i I , I , I I I , I , I

20 40 60 80 100 200 400 600 800 1000

Oscillation Frequency, Rz

Figure 26. Stability limits for planar oscillating inflow, NACA-8 compressor.

m o o -4

q4 & o

CD

f Fast Acting Bypass Door

~ ====_=====~L=_~,,,,~,, J L''

--Normal Oscillating Fuel Shock Wave Control Valve

/

a. With supersonic inlet in wind tunnel, Mach= 2.5 [52]

Test Cell Ducting

Primary . L ~ ~ ~ "j~

b. With airier system [53]

i

m O o

& O

Figure 27. Test arrangements for NASA J85-13 dynamic response experiments.

AEDC-T R-77-80

6 . 0

4 . 0

2 . 0

t

1 . 0 =o

0 . 6

0 . 4

0 . 3 4 . 0

D

Sym

D J 8 5 - 1 3 NASA E x p e r i m e n t w i t h S u p e r s o n i c I n l e t ( f r o m [ 5 2 ] ) N/,/SJ= 86.4% Mean P r e s s u r e R a t i o = 4 . 1 3

O J 8 5 - 1 3 Model C o m p u t a t i o n y/geeffi 87%, Mean P r e s s u r e R a t i o = 4 . 6

N o r m a l i z e d M a g n i t u d e = ( A P i / A P 5 ) f / ( A P i / A P 5 ) f f f i 1

f = 40 Hz

{ - - 4 I

S t a g e 2 61 i E n t r y [

- 3 5 7 C o m p r e s s o r 8 D i s c h a r g e

I I I I I I I I I I 6.0 8.0 10.0 12.0 14.0

D i s t a n c e f rom C o m p r e s s o r E n t r y P l a n e . i n .

I 16.0

a. Amplitude ratio

Figure 28. Variation of amplitude ratio and phase angle through the J85-13 compressor with oscillating compressor entry pressure at 40 Hz.

181

A E D C-T F1-77 -80

60 s~

C] J 8 5 - 1 3 NASA Exper iment w i t h S u p e r s o n i c I n l e t (from [52])

40 O J85-13 Model Computation

f = 4 0 Hz

"0

- 2 0

- 4 0 Stage 2 4 6 I Entry 8 [3

I Compressor lDischarge -eo I I I I I I I I I I I

4 6 8 I 0 12 14 16

D x s t a n c e f r o m C o m p r e s s o r E n t r y P l a n e , i n .

b. Phase angle

Figure 28. (Continued).

182

4.0 B Sym Mean Pressure Ratio

O0 t ~

2 . 0 - -

,P4

b~

1 . 0 --

t~ m

E 0 . 6 - -

0 Z

0 . 4 --

0 . 2 0 . 4

0

S o u r c e N L / 8 , %

J 8 5 - 1 3 NASA E x p e r i m e n t w i t h A i r j e t S y s t e m , { 5 3 ] 87

J 8 5 - 1 3 NASA E x p e r i m e n t w i t h S u p e r s o n i c I n l e t , L52] 8 6 . 4

J 8 5 - 1 3 Model C o m p u t a t i o n 87

4.13

4.6

•• ~ - ~ o o~. o- o o'-- "-Y "~- o o

-.-........._ o ...

N o t e : A m p l i t u d e R a t i o M a g n i t u d e N o r m a l i z e d t o V a l u e a t 1 H z

I I ,| I I I I I . I I I 1.0 2.0 4.0 6.0 10 20 40 60 100

Froquency, llz

Figure 29.

a. Stage 2 entry (amplitude ratio)

Dynamic response to oscillating compressor entry pressure, J85-13 compressor.

OO

v - 4

bD

®

e$

100

50 -

m

0 - -

m

- 5 0 - -

m

- 1 0 0

0 . 4

sym

O

J 8 5 - 1 3 NASA E x p e r i m e n t w i t h A i r j e t S y s t e m , [ 5 3 ]

J 8 5 - 1 3 NASA E x p e r i m e n t w i t h S u p e r s o n i c I n l e t , [ 5 2 ]

J 8 5 - 1 3 Model C o m p u t a t i o n s

N o t e : P h a s e A n g l e R e f e r e n c e d t o C o m p r e s s o r E n t r y P r e s s u r e O s c i l l a t i o n .

I I , , , I I I I I I I I J 1.0 2.0 4.0 6.0 10 20 40 60 100

F r e q u e n c y , i l z

b. Stage 2 entry (phase angle)

Figure 29. (Continued}.

oo

4.a

"O

e.4

B L4 O

4 . 0

2 . 0

J

I . 0 B

i

B

m

0 . 6 - -

0 . 4 --

0 . 2

0 . 4

Sym

0

.J85-13 NASA Exp+,rim#.nt with Airj~t System. [53]

J85-13 NASA Exl)(.rim,~nt with Sup(,rsorlic Inl,.t. [52]

, I 8 5 - 1 3 M o d e l C ~ m q ) u t a t i o n s

~'~=':----2..--:"- o o ,o., - o ^ • =..P ,a~'% # f ~w ."

i , I I I I I I I I I I I I 1 . 0 2 . 0 4 . 0 6 . 0 10 20 4 0

F r e q u e n c y , Hz

I ~ i

60 100

Co Stage 4 entry (amplitude ratio)

Figure 29. (Continued). ill

0

-,i

w 0

Q

I

hO

t=.4 b~

a

O,

1 0 0 - -

5 0 - -

0 - -

- 5 0 - -

- 1 0 0 0 . 4

s ~ - - - - - ' - - J 8 5 - 1 3 NASA E x p e r i m e n t w i t h

A i r j e t S y s t e m , [ 5 3 ]

. . . . . J 8 5 - 1 3 NASA E x p e r i m e n t w i t h S u p e r s o n i c I n l e t . [ 5 2 ]

0 J85-13 M o d e l Computation / /s~ %

0 0 0

I I I I I I i I I J I I l i i n i i l I . 0 2 . 0 4 . 0 6 . 0 1 0 2 0 4 0 6 0 1 0 0

F r e q u e n c y . H z

m O

~J & o

d. Stage 4 entry (phase angle)

Figure 29. (Continued).

Oo

.1

"el

N

0 Z

4 . 0

2.0 m

1.0 -- m

m

0.6 --

0.4 --

0.2

0 . 4

Sym

0

J85-13 NASA Exporiment with Alrj,Jt System, [53]

.185-13 NASA Experiment with Supersonic Inlet, [52]

J85-13 Model Computation

. o \ %

%

X X \

V ~t \

0

I I I I I I I I I I I I ' i ' t 1.0 2 0 4.0 6.0 10 20 40 60 100

Frequency, Xz

e. Compressor discharge (amplitude ratio)

Figure 29, (Continued), m

? M

OB

o

M

d 7. f:l

ffJ

1 0 0 -

50

-50

-100 0 .4

S_~vm "=------- J 8 5 - 1 3 NASA Experiment with

Airjet System, [53]

J85-13 NASA Experiment with S u p e r s o n i c I n l e t , [52] - - - - / ' /

0 J85 Model Computa t ion / I t

- - _ \

I 1 I I I I I III I I I I

1 .0 2 .0 4 . 0 6 . 0 10 20 40

F r e q u e n c y , Hz

I I I I I 6 0 1 0 0

f. Compressor discharge (phase angle)

Figure 29. (Continued}.

m o ? 3

& 0

A E D C-T R -77-80

6 . 0 ,

4 . 0

2 . 0

1 . 0

0 . 8

0 . 6

0.4

m

s,

N 0 . 2

E

0 Z

0.1

0.06

0.04

s ~ [3

0

- S t a g e 1 E n t r y

J85-13 NASA Experlment wlth S u p E r s o n i c I n l e t ( f r o m 1 5 2 ] ) N/4~ = 8 8 . 2 ~ , Mean P r e s s u r e R a t l o = 4 . 6 3

J 8 5 - 1 3 Model C o m p u t a t l o n N/4~ = 87~ , Mean P r e s s u r e R a t l o = 4 . 6 0

S t a g e 2 E n t r y

J

I I 4 7 8

f = 40 HZ

0.02

0.01" 0

Figure 30.

2.0 4.0 6.0 8.0 I0.0

Distance from Compressor Entry Plane, in.

12.0

a. Amplitude ratio

Variation of amplitude ratio and phase angle through the J85-13 compressor with oscillating engine fuel flow at 40 Hz.

189

AEDC.TR-77-80

6 0 -

s,m - - - 0 - - J 8 5 NASA E x p e r i m e n t w i t h

S u p e r s o n i c I n l e t , 152]

40 - --0-- J85-13 Model CompuSatlon

20 - f = 40 Nz

_2o t •40

l I I -60 0 2.0

S t a g e 1 E n t r y

I l I I I ] I I J I 4.0 6.0 8.0 10.0 12.0

D x s t a n c e f r o m C o m p r e s s o r E n t r y P l a n e , i n .

b. Phase angle

Figure 30. (Continued).

190

AE DC-T R-77-80

¢., ,o

4.1 . -4 ¢

¢ b;

.,..,

E L

c.

2 . 0

1 . 0

0 . 6 0

0 . 4 0

0.20

0. i0

0.06

0.04

0 . 0 2

0.01

0. 006

0.004

0.002 " 0 . 4

s_~ J85-13 NASA Experiment, [52]

O J85-13 Model, N/~ = Constant J 8 5 - 1 3 M o d e l , N / ~ = V a r i a b l e

D J85-13 Model. N//'O = Variable

~0

%,

P l = V a r l a b l e

1Hz A m p l i t u d e Ratio

&P5/AP13 = 0.16 AP5/AP13 = 0.0016 AP5/AP13 = 0.098 AP5/AP13 = 0.23

Aerodynamlc \ ~ Propagation, Rotor Speed and Pl Variation

"n \,,

aml c ~ • Propagation and~ ~ ' Rotor Speed ~ ~ Varlation ~ Z ~ ~ ~

0/0 Note All.Amplitude Ratios Normalized by 1 Hz Experi-

ZUpstream mental Data Value (0.16). Aerodynamlc Propagation Only i I I I I I l i l l l I l I ' l l l ] l 1.0 2.0 4.0 6.0 I0 20 60 lU0

Frequency, Rz

Re

Figure 31.

Compressor entry (amplitude ratio)

Dynamic response to oscillating fuel flow, J85-13 compressor.

191

m

~J

~q

m J:

100 --

50 -

0 --

D

-50 -

B

- 1 0 0 0 . 4

Sym

. . . . J 8 5 - 1 3 NASA Exper im ent , [52]

0 J 8 5 - 1 3 Model. N / , ~ = C o n s t a n t

A J 8 5 - 1 3 Model, N / ~ = V a r i a b l e

t:] J 8 5 - 1 3 X o d e l . N / / 8 = V a r i a b l e . " _ ~ Pl = V a r i a b l e

! i i I I i

1 . 0

b.

I I i n | | I I i i

2 . 0 4 . 0 6 . 0 tO 20

Frequency , Hz

Compressor entry (phase angle)

i I

4O

I I ! I m |

6 0 1 0 0

m

O n :m

~J

& o

Figure 31. (Continued).

~o

780

760

740

720

700

o 680

d = 660 , J

o 640 E

~" 69-0

o 600

580

560

540

50( 0 . 1 4

Figure 32.

I m p o s e d F o r w a r d B o u n d a r y C o n d i t i o n

/ / /

/ /

, T 1

\

/ / /

/ ~ ' L R.p Rate = A T = 250oR

/ / <i'r61 n i t i al 1

II ' i

E n g i n e E n t r y T o t a l T e m p e r a t u r e , T 5

! 3,600°R/sec

= 7 0 %

I

P~I IL~

O. 16 O. 18 0.20 0.22 0.24 0.26 0.28 0.30 0.32 Time, see

a. Inlet temperature

Rapid inlet temperature ramp to stall," NACA-8 compressor.

> m

o

& O

~D 4~

0.72

0.70

0.68

• o 0 . 6 6 oj

f., 0 . ~ 0 . 6 4 0

oj 4 J u w 0 . 6 2 k 1.4 0

¢J

O. 60

O. 58

\ / \ / \

/

!

> m O

9

R O

0.56

O. 14 O. 1 6 0 . 1 8 0 . 2 0 0 . 2 2 0 . 2 4 0 . 2 6 0 . 2 8 0 . 3 0 0 . 3 2

T i m e , s e c

b. Corrected rotor speed

Figure 32. (Continued}.

m ~D t~

O. 70

000

0 58 ~ ~ I =--, ~ n ~ " t , , - -

o.,t2 ~ rV b/ m tl V '

0.38 ~ ' ~ A ,

0.34

O. 30

0 . 1 4 0 . 1 6 0 . 1 8 0 . 2 0 0 . 2 2 0 . 2 4 0 . 2 6 0 . 2 8 0 . 3 0 0 . 3 2

Time, sec

Co Stage flow coefficients

Figure 32. (Continued).

m

m o

AE 0C-TR-77-80

240

O

200

°p4

4J

160

O

"~ 120

E 80

t,-

40

O

I n s t a b i l i t i e s

NACA-8, S t a b l e

NACA-8, Unstable

N / / 0 = 70%

0

~"~ ~-=-From [16]

S t a b l e (No S t a l l s )

0 L

0

Figure 33.

2 , 0 0 0 4 , 0 0 0 6 . 0 0 0 8 . 0 0 0 T e m p e r a t u r e Ris~ R a t e , °R/sec

1 0 , 0 0 0

Compressor instability caused by rapid inlet temperature ramps.

196

AE DC-T R-77-80

o , \ . Control Volume i , j , k I ~ I " ~ \ p Denotes + i / [ ~/~"? /~ ~ m D e n o t e d - 1 I I ! l ~ | I I ip = i + I, im = i - 1 ... ~ k ~ / ]

4 Control Volume (i,j,k) - (1,2.1) Front Vxew

i = 1 2

Side View

"jn=4

j -- 1

in = 5

%

i n - - 4

kn =6

Figure 34. Division of compressor into control volumes.

197

AEDC-T R-77-80

WC ,z j kp HCi j kp w i , j , k p

wz t j k HZ x j k

U xJ k

ARt J p k

• /

J -

WRi jp k

HRx jp k

v i Jp k

S ,

J

; /

y IWRIj k

L HRx3 k

v i i k AZ xj k

J

J

f

WZip J k

HZip J k

Uip j k

ACij k

wci j k

HCl J k

W i j k

r

Figure 35. Control volume velocities and fluxes.

198

AEDC-TR-77-80

Volume i j k

WC

Cont Volume / ~X ~ I / ~ / / i j kp ~

Flow

R o t o r Gap S t a t o r

/ P S h i g ~ .ic

/

Orifice Flow Analogy

m m

PShig h

WC

PSIo w

Ro t a t ion

Figure 36. Circumferential crossflow.

199

AE DC-T R-77-80

C o n t r Volum i 3P

C o n t r Volum i j k

Blade Lead ing Edge

, ~ e c o n a a r y F l o w i n B l a d e Passages

w

Figure 37, Radial crossflow.

200

AE 0C-TR-77-80

I n d u c e d R a d i a l

F low

R a d l a l l y DI ow

U n i f o r m F l o w

U 0 *

V e l o c l t y T r i a n g l e s f o r C a s e A

Figure 38. Radial distortion effects.

201

AE DC-TR-77-80

p

0 0 3600

90 ° 270 °

f 0 iO °

U $

e = 90 °

f

U | ~ Vswirl

180 ° e i IS0£ U $ e]27o

Z J IVswirl ° U

a. Steady flow effects

F~gure 39. Circumferential distortion effects.

202

AEDC-TR-77-80

270 . . . .

9 0 ~ . / / / P l:ov,' " " ,%_ / P

I

~"'- ~ dy.

0 360 5, deg

0 360 0, deg

~dyn

b. Unsteady flow effects

Figure 39. (Continued).

203

AE DC-T R-77-80

C~

/ i | ~ C i s t a l l

' f

V C d~ re f

Angle of Attack, a

m = 40 1.0-- -- / m= 20

. ~ . g e of Exper l - / ~ -<> - I / " _ m e n t a l Data, ~ ~ , ~ ~

08 t361 --~ /~.~\ h-'l

o 0 0.01 0.02 0.03 0.04 0.05

C ~a Vre f %t ' r ad lans

c. Unsteady stall effects

Figure 39. (Continued).

204

AE DC-T R-77-80

Ploc/Pavg Pattern Amplitude, Pmax - Pmin

= lO.2% Pavg

1.00 - 1.02

90-- -- 270

1.00 - 1.02

- 1

9 4 - 0 . 9 6

Figure 40.

0 . 9 3 - 0 . 9 4

/ i / 180 s Range of

Ploc/Pavg 0.96-0.98 in Region

Indicated

Compressor entry total pressure distribution for combined distortion pattern.

205

AEDC-TR-?7.80

' ~ P l o c / P a v g

1.018 "--~

1. 029 - - ~

1. 018 -.-.

1 .018

1.018

.018

.029

.018

018

1. 024

1. 024

1 .024

1.024

1.024

k = l

2

1 .002

0.9595

3 4

0 . 9 9 1 6

O. 9916

O. 9916

1 . 0 0 8

1 . 0 1 8

Figure 41.

0 . 9 3 7 0

0 . 9 3 7 0

0 . 9 3 7 0

Pmax - Pmin = 0.092 ~ ~ - ' - 0 . 9 5 4 1 P a v g

0.9884

Compressor entry pressure input values for combined distortion pattern.

206

A E DC-T R-77-80

O --d 4~

O

k C

E O

4.0

3.5

3.0

2.5 31

1.3% ~

~S ~,,a~ ~ ~

f 5.3%

O. 8%

Sym

O Experimental Measurements with Distortion, [43]

Experzmental Stability Limit with Distortion. [43]

Model Computations with Distortion

No Distortion

i I I I I 1 32 33 34

Corrected A i r f l o w - - = - - ibm/sec

o',

I = • 35 36

~ / , ~ = 9 0 %

Figure 42. Comparison of computed and experimental stability limits with combined distortion pattern, XC-I compressor.

207

AE DC-TR-77-80

0 Model C o m p u t a t i o n s

0 Expe rzmen ta l 1 .025 F Measurements , [43]

!

0 . 9 7 5 L D i s c h a r g e

L

o

i .025

1. 000

0 .975

1 .025

1. 000

I 0.975

1.025

1. 000

0.975

I. 050

1. 000

O. 950

Figure 43.

~ " 4-~ ====~>'0~ "'~-- G "FI

- ' , , , ' 7 - - ~,N .~./ .<>- .... o---~-.. ~'-~o 6~ ..'~--- "

-

- " ~ - o--.d' ' I I I I I I I I I I I

0 60 120 180 240 300 360

C i r c u m f e r e n t i a l Location, 6. deg

a. Total pressure

Circumferential profiles in tip region with combined distortion pattern, XC-I compressor.

208

AE DC-T R-77-80

1 . 0 2 5

1. 000

0 . 9 7 5

1.025

1. 000

0.975

:) 1.025 a

bo

1. 000

u

o 0. 975

1. 025

1. 000

0.975

1 . 0 2 5

1. 000

0 . 9 7 5

s_~ O Model C o m p u t a t I o n s

[3 E x p e r i m e n t a l M e a s u r e m e n t s , [43]

/ Discharge / O " ~ " ' 0 ~" 6 I " - - 6

/ D - D ~ o

o o - - d... ~ ' ~ "o -cu

.0"-- -0. "

. . . . [] _ . -'0 ~"'01 ":OO" []

- / O . O " -O, ',

_ 2 ~ O ~ _ _~/. D ~ O ~ ~ 1 3 . D "--:8--o- - -

- - S t a g e 1 P r e s s u r e ~ E n t r y Zone

----~0 . . . . 0 =--~D - - ~ O D - ' ~ D ' = O ---~- D ---D O=D =~D---O--D:D

l I I I i I I I I I i • 0 60 120 180 240 300 360

C i r c u m f e r e n t i a l L o c a t i o n , O, d e g

b. Total temperature

Figure 43. (Continued).

2 0 9

AE DC-TR-77-80

~a "w 5- I

!

K

[] 0 -

10--

m

= m

[]

!

M

E

n

0 -2

Compressor

Discha rge

Stage 1 i !

,, m ~ ..0~ ~ --

/~O ~ O ~ O ~ o

O O

O - - _ .

O O

Model Computa t ions E x p e r i m e n t a l Measurements from [43]

" -0 O ~ •

~ o q ~ ~ • XC-i Compressor ~ ~]

Experiment Mode___1 ~ %

PR 3.71 3.72

3 3 . 8 9 3 4 . 0 5

I I I I i I # I I 0 2 4 6 8 Distance into Compressor, in.

a. Tip region

Figure 44. Comparison of computed and measured circumfer- ential distortion attenuation for combined pattern.

210

AE D C - T R - 7 7 - 8 0

C o m p r e s s o r D i s c h a r g e

S t a g e I E n t r y 3

. 5

/ O" ---- O /

j ,a

o ==

10

B

"u -

B

"" O ~ D

!

a _ N

0 ! - 2

r i m

%%%%

Model C o m p u t a t i o n s

E x p e r i m e n t a l Me a sur e ment s from 1 4 3 ] %%

0 ,,

! I I I I l I I i 0 9. 4 6 8

Distance in to CompreBsor. in.

b . Mid-span region

Figure 44. (Continued).

211

AEDC-TR-7740

!

E E-

e-

. p 4

E

I

X

E

5 i

m

i

i

i

0 -

1 0 -

m

m

n

m

5 - -

m

n

0 - 2

Stage 1 E n t r y

s

/

- - 0

/ /

3 4

Compressor D i s c h a r g e

"[3

/ o

_o

sym

O

D

Model Computations

Experimental Measurements from [43]

o

/ -Q /

\ \ \

\\

l l I I I i I I J

0 2 4 6 8 D;stance into Compressor, in.

c. Hub region

Figure 44. (Continued).

212

t~

8 . 0

7.0

"4

6 . o t~

5.0

m

4.0 m

_ Compressor St.',1~_e 1 hntry 2 _3 4_ Discharge

- ~',I /3 ? " ° " ' , /, \

O D I 0.98 1.0 1.02

0 , 9 8 1 . 0 1 , 0 2 I l I

• / 0 O. . 1 .0 1.02 1.04

I I I [0= 210 deg] 1.0 1.02 1.04

Ploc/Pradial avg

[] 0 \ ~ t I .96 0.98

L , I I I Sym

o . 9 8 t.O 1 . 0 2 1.04 *.06 0 rn

Model Computations

ExperLmental Measurements, [43]

a. Total pressure

Figure 45. Radial profiles with combined distortion pattern. m 0

9 -4 =m

"4

0

m

4~

8.0

7.0

6.0

5.0

4.0

S.....~ta.~e 1 Entry

c~

I i I 0.98 1.0 1.02

Compressor 2 3 4 Dischar6e _ _ w

j o o., ,o.o. A I I I I I I

0 .94 0 .96 0 . 9 8 1 .0 1 .02 1 . 0 4

I I l I I 0 . 9 6 0 . 9 8 1 . 0 1 .02 1 .04

[ I I ] J [ 6 = 210 d e g ] 0.96 0.98 1 0 1.02 1.04

" T l o c / T r a d i a l a v g

sym

0

0

M o d e l C o m p u t a t i o n s

Experimental Measurements, [43]

b. Total temperature

m

n

3

& o

Figure 45. (Continued).

10

8 . 0

7 . 0

-s

" 6 0 --

5.0 -

w

4.0 -

St~go 1 Entry

o Compressor

2 3 4 D i scha rge

t tD ~ ~ ED C~_ L

,I ",;

/ / ./~- ~" CO 0.98 1 . 0 1 .02

L [D 0 . 9 8 1 . 0 1 .02 r • i i i i

d~ 0.96 0.98 1.0 1.02 1.04

L I I I I 0.96 0.98 1.0 1.09. 1.04 PloclPradial avg

L I I I I 0 . 9 8 1 , 0 1 .02 1.04 1 .06

s~m

O

B Model Computa t ions

E x p e r i m e n t a l Measurements , {43]

a. Total pressure

Figure 46. Radial profiles with uniform entry flow. m

O O

& o

t , o

R . O -

7 . 0 -

c-,

,=4

, / )

6 . o - c~

m

5 . 0 -

m

• ! . 0 - -

s'ta e_~_/ L.nt~

I

)

I I I I I I

)

I I I 0 . 9 8 1 . 0 1 . 0 2

Compressor 2 3 4 D i s c h a r g e

6 Y , , ~ 6~ 0 . 9 6 0 . 9 8 1 . 0 1.02 1 .04

Ii , - F i , , J ,/V" m 0.94 0.96 0 .98 1.0 1.02 1.04 I,* I i I ' J

( ~ 0 . 9 6 0 . 9 8 1 . 0 1.02 1.04

I I I , ' •

0 . 9 6 0 . 9 8 1 .0 1 .02 1 . 0 4 T l o c / T r a d i a l a v g

s_ym

0

El

Model Computatlons

Experimental Measurements, [43]

b. Total temperature

I 1 . 0 6

m

O f~

m O

Figure 46. (Continued).

AEDC-TR-77-80

Figure 47.

\ !

~ 1 . 1 6 - 1.20

. . t . . . ~ l ~ l . 08 1.16

O I

~ 1 . 0 4 1.08 1

~ ~ 1 . o 2 - 1 . o 4

1.01 - 1.02 1.00 - 1.01

~/~stall S

a. Stage 1

Flow coefficient distribution just prior to instability with combined distortion pattern.

217

A EDC-TR-77-80

1.16 - 1.20

.08 - 1.16

1.00 - 1.02

1 04 - 1.08

/ / L S t a l l e d

1 . 0 2 - 1 . 0 4

b. Stage 2

Figure 47. (Continued).

. 98 - 1.00

218

AEDC-TR-77-80

--- 1.16 - 1.20

!

f 1.08 - 1.16

1.04 - 1.08

1.02 - 1.04

c. Stage 3

Figure 47. (Continued).

219

AE DC-TR-77-80

1.04 - 1.08

- 1.04

1.00 - 1.02

- - 0 . 9 8 - 1.00

S t a l l e d

0.94 - 0.98

d. Stage 4

Figure 47. (Continued).

220

AEDC-TR-77-80

Location

Total Pressure n I m

R

l n

-- I =

l I I,i,I

V v 1 r v '1

I ~ ~ ~ ~ I ii2,1 [ ~ l 1 , 3 , 1

m m [

_ I _ I

: I : I

l l ~ l i,j,k I 1 , 6 , 5

t = t 1 t --. t 2 t = t 3 t -- t 4

I / I I i

/,

Figure 48. Time-variant distortion.

221

I J I J I - )

0) >

,F4

r ~

1 . 0

0 . 8

0 . 6

0 . 4

0 . 2

0 2 4 6 10 20 4 0 6 0 100 2 0 0 4 0 0

F r e q u e n c y , Hz

n

co o

Figure 49. Relative frequency spectrum for time-variant distortion pattern.

AEDC-TR-77-80

1.0 F ~ ~Range of Experimental

0 . 9 ~ [71

O~ "~0 ~ o.8 I I I I I

0 1 2 3 4 5 G

~ rms Amplitude, AP /P, % m ~ rms

= a. Correlation with fluctuation amplitude

• ~o ~ .r'4 4-) ~ 0 4 Steady-State

Dis tor t ion ~ l o O Time Variant . ,~

= I__ ~ Distor t ion

c a ~

~. . 4 J

o . 9 - v~ 2% r m s ~

0 0 . 8 -- 4% rms--/

0.7 I I I I I 10 20 30 40 50

Instantaneous Distortion Pattern Magnitude, (Pmax - Pmin)/Pavg' %

b. Correlation with instantaneous distortion magnitude

Figure 50. Loss in stability limit pressure ratio with time-variant distortion

223

AE D C-T F1-77-80

4 . 0

2

3 . 5

¢ .

O

&9

t~ E O

3 . 0

2.5

,jpS

- ' " ")---No .. ~ 4~. _. . ~ Distortion

I0,~ %

-

N / ~ = 90% 20.~

T i p

e~ ee

1 . 0

l o c / P a v g

O 5%-Distortion

10~

U 20.%

O A tl I I \ \ J

1 . ! Hub

0 . 9

i I I I I I I I I I 31 32 33 34 35 36

C o r r e c t e d A i r f l o w , 6 ' I b m / s e c

a. Tip radial distortion

Figure 51. Computed radial distortion influences on stability.

224

AEDC-T R-77-80

0

O

E O

-- " " D i s t o r t ion

4 . 0 _ ~ ~

3,5 - -

m

3.0 -

m

2.5 31

5%

10% 2o~ ~ ,

N / ~ = 9 0 %

Tip

Hub 0 . 9 1 .0 i . I

P l o c / P a v g

i i I [ I 32 33 34

Corrected Airflow, 6 '

O 5%-Distortion

n 1o%

D 20,~,

I I I I 35 36

I b m / s e c

b. Hub radial distortion

Figure 51. (Continued).

225

AE DC-TR-77-80

1.1

1.0

s Range of J85-13 Engine Test Data for Flow Range from 77 to 98%, from [16]

y Model Computa t ions f o r XC-1 Compressor , N / ~ = 9 0 % w_q =

I I I I I 5 10 15 20 25

D i s t o r t i o n P a t t e r n Magni tude ,

• ,..4 C 4 J r = . er.,., 0 . 9 - - 1

tw :3, '~

¢~,,.C t . . ,~ ~, .,, 0.8

• ,-.I 0

'~ ~ ( Pmax Pmin)/Pavg, %

~ 1.11-- a, Tip radial distortion

1 ~ | ~ Range of J85-13 ~ I ~ s" Tes t Data

~o ~ ~ /---- Model Computations,

0.9 _ ~,~ _/ XC-I C o m p r e s s o r

0 . 8 I I I I I 0 5 I0 15 20 25

Distortion Pattern Magnitude, (Pmax - Pmin) /Pavg" %

b. Hub radial distortion

Figure 52. Comparison of radial distortion model compu- tations with experimental results.

226

AE DC-TR-77-80

0 ..4 4~

~O W Q;

fw o (0

fw

E 0

U

4 . 0

3 . 5

3.0

m

2.5 31

B ~ e e l s

S S

S ~ e e ~ ~

~ ' " ~ ~ - No

N / / e = 90%

Sym

PH 0 £

0

0

I I i I i I 32 33 34

C o r r e c t e d A i r f l o w , l b m / s e c

PH - PL

P avg

5~

I0~

15%

20%

I I I 35 36

Figure 53. Model computed influence of circumferential pressure distortion on stability, XC-1 compressor.

_~7

AE DC-T R-77-80

1. O0 ~ ~". - . y Range o f

~ ~ / ~ / ~ / / ' ~ ~ ~ E x p e r i m e n t a l D a t a

0.

85 / -- Model • ~ o ~ C o m p u t a t i o n s ,-3 ~1

.,-i

.O ~0 O

• ~ g., r n ~

O . , . d [] O . , ' d

O

0 . 8 0 --

O. 75 -

0 . 7 0 -

0 . 6 5

0

s_~

0

D

0

A

Model Computations, XC-I Compressor,

Wv~" = 90% N / ¢ ~ = 90%,

4 - S t a g e C o m p r e s s o r ,

= lO1%, [16] 6

5 - S t a g e Compressor,

w4~ 6 = 93%, [ 1 6 ]

J 8 5 - 1 3 C o m p r e s s o r , [31]

I I I I 5 i 0 15 20

Distortion Pattern Amplitude, (PH - P L ) / P a v g ' %

! 25

Figure 54, Comparison of circumferential distortion model computations with experimental results.

228

AEDC-TR-77-80

O

E o o o o

W-

re O

O~

o

Z~

0

0

1 . 2 -

1 .1

1 . 0

0 . 9

-2

D a t a f o r 10~, 6 0 - d e g P a t t e r n

Stage 1 Exit

2

C o m p r e s s o r D i s c b a r g e

o.-- t "I-o_ . . . . . . ~ . . . .

Stage 1 / / _ ~.....O"-'-'<2 Entry / Z~"

- I . . ) " Y , s _ ~

• 0 Determined from ~• / /\ 0 Determined • • S , s - / ~ Experiment, [43]

~ ~ \ O Model Computed D i s t r i b u t i o n , /

CXFC = O. 6 D i s t r i b u t i o n U s e d - - - J f o r Combined Zl Mode] Computed D i s t o r t i o n P a t t e r n D i s t r i b u t i o n , Case CXFC = 0 . 6

_ S t a g e 1 C r o s s f l o w Area D o u b l e d

i I I I l I I I I I

0 2 4 6 8

D i s t a n c e i n t o C o m p r e s s o r , i n .

a. Circumferential crossflow distribution

Figure 55. Circumferential crossflow magnitudes and effects.

229

AE DC-TR-77-80

0

0

E 0 U

4.0

3.5

3 . 0 m

m

m

2.5

31

o

s~ ~ /Stage i Crossflow

_" ~ t...oa ,~. " f " ° L Cr°ss fl°w---~ 2 , 1.0%

o • 1.4% ~ ~ •

O,

d,

N/J6 = 90%

Svm _.x..__

O

A

Crossflow. CXFC = 0.6, 10t, 60-deg Distortlon

Stage 1Crossflow Doubled, 10~. 60-deg Distortion

No Distortxon

I I I I I I 32 33 34

Corrected Airflow

b. Effect on stability limit

I I 1 35 36

Figure 55. (Continued).

230

AE DC-TR-77-80

_ \ - ~ ~ ~ N _ _ No

180 °

~ 3 .5

~ L 3 ° ° " ~ - - PH - PL = I 0 % ~ i Pavg

3 . 0

2 . 5 ' - 31

o w o

[] o o

PH

O

D

L

Low P r e s s u r e Region Angu la r

Ex ten t~ 0-

30 deg

60

180

32 33 34 35 36

w,~ C o r r e c t e d A i r f l o w , --~--, lbm/sec

N / / ~ = 90~

Figure 56. Model computed influence of circumferential distortion angular extent on stability, XC-I compressor.

231

AE DC-T R-77-80

I O0

0 9 8

O R "~ O ~ 0 9 6

k ~ 0 9 4

°F,,I 0

~ 0 9 2

.,..i m

m ~

o F 0 9 0 o ~ • .4 ,.] 4~ ~ o

0.88

0.86

-

-- I Pmax - Pmln = 10~ [

I P a v g I - s _ ~ m_

O J85-13 E x p e r i m e n t a l D a t a , [ 4 ]

- O I

10

- h 20

O 40

- r~ 1 8 0 - d e e D i s t o r t i o n

i I I I I I I I I I 20 40 60 80 I00

D l s t o r t L o n P a t t e r n A n g u l a r E x t e n t , d e g

Figure 57. Effect of unsteady airfoil stall properties on stability margin.

232

A E DC -T R-7 7 -80

~..... ~ " " ~'*" ° " " ~ i0 Distortion 4.0 ....~

.~ . j , . . . ~ - - % o , , , , , / - - . , ~ -- , o~

= ~ o ~ - / ~ ~ -N-\o',

I"1 15%

- 0 2o~

M T H

8 9 . 7 % 89. 370

89.1%

88.9%

m

2.5 I i I I I I I I I I 31 32 33 34 35 36

Corrected Airflow, W/O Ibm/sec 6 '

Figure 58. Model computed influence of circumferential temperature distortion on stability, XC-I compressor.

233

AE DC-T R-77-80

1.00

0 . 9 5

o 3 • ,~ O 4 J , - 4

0 . 9 0 -

~ . , ~

0.85 -- .J

• ~ O

~ 0.80 -

C'~ E

O -~ O. 75 --

¢d O

0 . 7 0 --

0 . 6 5 0

Figure 59.

Range o f E x p e r i m e n t a l D a t a

o M o d e l • C o m p u t a t i o n s

~ J

o M o d e l C o m p u t a t i o n s , XC-1 C o m p r e s s o r ,

N / ~ = 90%

[3

<>

C o r r e l a t i o n B a s e d o n F o u r C o m p r e s s o r s , [ 1 6 ]

J 8 5 - 1 3 , N/CrO = 87%, [9 ]

J 8 5 - 1 3 , N / / 9 = 94%, [ 9 ]

[]

I l I I I 5 10 15 20 25

D i s t o r t i o n P a t t e r n A m p l i t u d e , (T H - T L ) / T a v g , %

Comparison of circumferential temperature distortion model computations with experimental results.

\

234

AE DC-TR-77-80

APPENDIX B

COMPRESSOR DETAILS AND STAGE CHARACTERISTICS

Allison XC-I Compressor

Figure B-I (Appendix C) 1 shows the overall com-

pressor map for the Allison XC-I lift engine, four-stage

compressor. Key dimensions and design performance are given

in Table B-I (Appendix C). 1 Stage characteristics used in !

the model are given in Figs. B-2a through B-2d. The mean of

the outer and inner radii was used in the stage flow and

pressure coefficient calculations. All the information was

taken from [43]. Dimensions and data not tabulated in the

report were obtained from scaling drawings and figures.

NACA-8 Compressor

Figure B-3 gives the overall performance of the

NACA eight-stage research compressor. Principal dimensions

and design performance are given in Table B-2. Stage

characteristics are given in Figs. B-4a through B-4h. Mean

radii were used in the flow coefficient, equivalent pressure,

and equivalent temperature ratio computations. All infor-

mation on the NACA-8 compressor was taken from [44, 45, 46].

IAII figures and tables for Appendix B appear in Appendix C.

235

AE DC-TR-77-80

General Electric J85-13 Compressor

Overall performance of the J85-13 compressor

modeled during this study is shown in Fig. B-5. The

variable guide vane and compressor bleed schedules used in

the model (and for which the compressor map of Fig. B-5

applies) are given in Fig. B-6. Key dimensions and design

point performance is given in Table B-3. Rotor inlet air

angles for which the stage characteristics were calculated

are also given in Table B-3. Stage characteristics are

shown in Figs. B-7a through B-7h. The first-stage flow and

pressure coefficients were re-defined for this work in terms

of a semi-empirical modified flow and pressure coefficient

to account for variable guide vane effects. The modified

coefficients are,

and

cos Sm= 1 (B-l)

--- sin 3

P r ) l,i

= , (B-2)

3 2 4m [ 1 - ~m sin "i]

where ¢ and ~P follow the conventional definitions given in

the text. The value of 3 used is the guide vane angle from

Fig. B-6.

Information on ~he J~5-13 was taken from [47].

236

AE DC-TR-77-80

APPENDIX C

FIGURES AND TABLES FOR APPENDIX B

237

6.0

Compressor S t a b i l i t y Limit

o 5.0--

==

4 . 0 - -

r.4 p4

0 > 3 . 0 - - o

r,,., o

E 2.0-- o

1.0

20

AE DC-T R-77-80

5 90 100

C o r r e c t e d R o t o r S p e e d , P e r c e n t D e s i g n

I I I I 25 30 35 40

we~- Corrected Airflow, 6 , Ibm/sec

Figure B-I. XC-I overall compressor performance characteristics.

238

AE DC-TR-77-80

0 . 9 ~

~J

® 0.8 I °~1

q-I q~

0 . 7

0 . 5

0 . 4 --

~ 0 . 3 - -

0

i,t

• 0 . 2 - N

0 . 1 0 . 2

I I I I 0.3 0.4 0.5 0.6

Flow Coefficient,

a. Stage 1

Figure B-2. XC-I compressor stage characteristics.

239

A E D C - T R - 7 7 - 8 0

0 . 8 .,=1

0 , 9

0 . 7 I I I I

0 . 5

0 . 4

4~

k~ m ¢ 0 . 3 O

b.i

0 . 2 - L. e'

0 . 1 0 . 2

l I I 0.3 0.4 0.5

Flow Coefficient, ¢.

b. Stage 2

Figure B-2. (Continued).

I

0.6

240

AE DC-TR-77-80

0 . 9

C"

0 . 8 (D

- 4 %.4

0 . 7 I I

0 . 5

0.4

-4

°~

0.3 --

C

¢ 0 2 - -

0 . 1

0 . 2

I I I 0.3 0.4 0.5

Flow Coefficient,

c. Stage 3

Figure B-2. (Continued).

I 0 . 6

241

AE 0C-TR-77-80

0.9

0.7

= 0.8

-,d

-,d

0.5

0.1 0.2

0.4 --

4J

0 .rd b.4

0.3 -- OJ C 0

k

• 0.2 -- C.

I I i 0.3 0.4 0.5

Flow Coefficient, ¢

d. Stage 4

Figure B-2. (Continued).

I 0 . 6

242

AEDC-T R-77-80

ii

i0

9

= 7

~ 6

~ 5

~ 4

~ 3

2

1

0

C o r r e c t e d A i r f l o w .

!

!ii!!- o ~o ~o~°

I i0 20 30 40 50 60 70

w/E~ 6 , Ibm/sec

Figure B-3. NACA-8 overall compressor performance characteristics.

243

AE DC-T R-77-80

1 . 2 8 -

~ . ~ 1 . 2 4 -

~ n e i. 20 -

~ m 1.16 -

eE 4~ m ~ 1 . 1 2 -

1.08 ] l

1.5

o

~ 1.4 m

m z ~k ,.u m

~o ~ 1.3

1.2 I I i i I 0.2 0.3 0.4 0.5 0.6 0.7

Flow Coefficient, ¢

a. Stage 1

Figure B-4. NACA-8 compressor stage characteristics.

244

1 . 2 0

1.18

1.08

O

~ 1 . 1 6

g ~ 1 .14

E

1 . 1 0

I I I [

AE DC-TR-77-80

.Ill m

>he

& ~ Z

~ 2 . . .

1 . 4

1 . 3

1 . 2

0 . 2

I I I I I 0 . 3 0 . 4 0 . 5 0 . 6 0 . 7

F low Coefficient, ¢

b. Stage 2

Figure B-4. (Continued).

245

A E OC -T R -77-80

1.12

~I. i0

=~1 08

~1 06 ~E " m[..,

I . 04 I I I I I

~ 0 ~.~

.4

1 .5

1 .4

1 , 3 m

1.2 0 . 2

I I I I 0 . 3 0 . 4 0 . 5 0 . 6

F l o w C o e f f i c i e n t ,

I 0.7

c. Stage 3

Figure B-4. (Continued).

246

AE DC-T R-77-80

O

4J

~E

I. 12

i. I0

1.08

1.06 I I I I

+~

O

cO

1.5

1.4

1.3

1.2 0.3

I I I 0.4 0.5 0.6

Flow Coefficient,

d. Stage 4

Figure B-4. (Continued).

I 0 . 7

¢

247

AEDC-TR-77-80

wo 1.12

~ 1 . 1 o

, a ~ 1 . 0 8 g~ , oE rn~, 1 .06

=0

e=q ,~

M

5'}

1.5

1.4

1.3 --

1.2

0.3

I I I 0 . 4 0 . 5 0 . 6

Flow C o e f f i c i e n t , ¢

I 0.7

e. Stage 5

Figure B-4. (Continued).

248

AEDC-TR-77-80

~o 1.12

1.10

1.08

1.06 r i I I

=0

cr~

1.5

1.4

1.3

I

1.2

0.3

I J I

0 . 4 0 . 5 0 . 6

Flow C o e f f i c i e n t ,

I

0.7

f. Stage 6

Figure B-4. (Continued).

249

AEDC-T R-77-80

~o 1.10

r-4

~ 1.08

~ • 1.06

m ~ 1.04

4~

= 0 ~J.~

"4

1.4

1.3

1.2

1.1

1.0 0.4 0 .5 0 .6 0 . 7

Flow C o e f f i c i e n t .

I

0.8

g. Stage 7

Figure B-4. (Continued).

250

AE DC-TR-77-80

i. I0

~J

1 . 0 8

1 . 0 6

1.04

1.02

1.00

1.3

1 . 2

1 , 1 m

1 . 0 - -

0 . 9 --

0 . 8

0 . 4

I 0 . 5

~C

I 0 . 6 0 . 7 0 . 8

Flow C o e f f i c i e n t , ¢

h. Stage 8

Figure B-4. (Continued).

251

t~ t~

8

~:'0~ 7 ~ 0 / Experimental

o 5 ~ / ! or

8 4

• 8

325 30 35 40 45

W~rO Ibm/sec Corrected Airflow, ~ ,

m

? -4

& O

Figure B-5. J85-13 overall compressor performance characteristics.

A EDC-TR-77-80

40 - ' 0

d "~ 30 -

e~

¢ 2 0 - - ¢1

1O-

F-4

-10 50

2 . 0

o

" 1 . 5

1.0 -

0.5

° I (¢/ -

I~ -

E 0 -

0 • 7 5

Vane Axial

I I I I I, 60 70 80 90 100

N/~-, percent

S t a g e 4

S t a g e 5

S t a g e 3

80 85 90 95

N/VrS, pe rcen t

100

Figure B-6. Inlet guide vane and compressor bleed

schedules for J85-13 model.

253

0 . 9 0

0.85

0

.,.4 ~J

"~ 0 8 0 ~ •

0.73

0.3

~%~ 0 .2

e -

°r,d

~.~ ~ , . ~ 0 . 1

0

AE DC-T R-77-80

l I I I F l o w C o e f f i c i e n t ,

0.3 0.4 0.5 0.6

Modified Flow Coefficient, Cm

a. Stage 1

Figure B-7.

0.7

J85-13 compressor stage characteristics.

254

0 . 9 0

0.85

=

w~ ,~ O. 8 0

O. 76 I I

AEDC-TR-77-80

4a

-F4

0.3

0.2

0.1

0.3

I I I 0.4 0.5 0.6

Flow Coefficient,

b. Stage 2

Figure B-7. (Continued).

I 0.7

255

AE DC-TR-77-80

0.9

= 0.85

0 . 8 I I I I

C~

°~

-I

O~

0 . 4

0 . 3

0 . 2 m

O . 1 0 . 3

I I I 0 . 4 0 . 5 0 . 6

F l o w C o e f f i c i e n t . ¢

I 0.7

c. Stage 3

Figure B-7. (Continued).

256

AE 0C-TR-77-80

0 . 9 0 -

" 0 . 8 5 -

O. 8 0 I I I I

-r4

O

O

0.3

0.2

0.1

0 0.3

I I I 0 . 4 0 . 5 0 . 6

F l o w C o e f f i c i e n t .

I 0.7

d. Stage 4

Figure B-7. (Continued).

257

AEDC-TR-77-80

O. 9 0 -

= 0.85

£ 0

0

~ 0 . 8 0

0.76 I I i I

0.3

4~

0 . 2 ";" .t4

0

0.1 --

0 0.3

I I I 0.4 0.5 0.6

Flow Coefficient,

I 0.7

e. Stage 5

Figure B-7. (Continued).

258

0 . 9 0 --

0.85

%.

0.80

0.76

0 , 3 m

AE DC-TR-77-80

l i l l

q. , e .

. , .q ° .

o,..4

L .

Y .

L ' L_

e ~

0 . 2 m

0 . I - -

0

0 . 3

I 0 . 4

l I 0 . 5 0 . 6

F l o w C o e f f i c i e n t ,

l 0.7

f. Stage 6

Figure B-7. (Continued).

259

AEDC-TR-77-80

0 . 9 0

O. 85

0.80

0 . 7 6 I I I I

.--D-

.-4 U

0

0 . 3

0 . 2

0 . i

0 0 . 3

i I I 0 . 4 0 . 5 0 . 6

F l o w C o e f f i c i e n t , ¢

l 0.7

g. Stage 7

Figure B-7. (Continued).

260

O. 90

0 . 8 5

~J e~

.=4 0

;2 ~ O. 8 0

O. 7 6

0 . 3

I I

9 -

p,

.-I

O

k

0 . 2

O . 1 B

0

0 . 3

I I

A E D C - T R - 7 7 - 8 0

I 0 . 6 0 . 4 0 . 5

Flow C o e f f i c i e n t ,

I 0.7

h. Stage 8

Figure B-7. (Continued).

261

AE DC-TR-77-80

Table B-I. Detroit Diesel Allison XC-I Four-Stage Lift Engine Compressor Model Details

Geometr[

Stage Area Length Outer

(Station) (ft 2) (ft) (ft)

Radius Inner

(ft)

1 1.136 0.2165 0.6991 0.3565

2 0.8173 0.1558 0.6903 0.4651

3 0.6610 0.1558 0.6903 0.5159

4 0.5045 0.1386 0.6838 0.5540

5 0.4026 - 0.6816 0.5800

Design Point Performance

N = 18,030 rpm

W~'{i = 38.7 ibm/sec

6

PR = 4.92 (maximum)

262

A EDC-T R-77-80

Table B-2. NACA Eight-Stage Research Compressor Details

Geometry

Location Area Length Outer Inner

Station Stage (ft 2) (ft) (ft) (ft)

1 * 2.182 2 * 2.182 3 * 2.182 4 * 1.679 5 1 1.679 6 2 1.311 7 3 1.173 8 4 0.9960 9 5 0.8352

10 6 0.6860 ii 7 0.5681 12 8 0.4737 13 ** 0.4737 14 ** 0.3129 15 ** 0.3129 16 ** Variable

0.475 0.475 0.475 0.475 0.475 0.450 0.283 0.292 0.258 0.242 0.250 0.242 0.242 0.500 0.500

0.8333

0.8333

0 0 0

0.4000 0.4000, 0.5265 0.5667 0.6100 0.6546 0.6900 0.7167 0.7373 0.7373 0.7712 0.7712

Inlet duct.

Burner.

Design Point Performance

N = 13,380 rpm

- 65 ibm/sec

PR = I0.0 (maximum)

263

AE DC-TR-77-80

Table B-3. General Electric J85-13 Eight-Stage Compressor Details

Location

Station Stage

Geometry Radius

Area Length Outer Inner

(ft 2) (ft) (ft) (ft)

Rotor Inlet

Air Angle (deg)

1 *

2 * 3 * 4 * 5 1 6 2 7 3 8 4 9 5

10 6 11 7 12 8 13 ** 14 ** 15 ** 16 **

1.300 0.1856 1.300 0.1856 1.300 0.1856

1.150 0.1856 1.066 0.1856 0.8395 0.1364 0.6821 0.1116 0.5630 0.09767 0.4690 0.08725 0.4055 0.07892 0.3698 0.07750 0.3580 0.09883 0.3542 0.i000 1.312 0.i000 0.3542 0.1000

Variable

0.6708 0.6708 0.6708 0.6708 0.6458

0.6458

0 1900 0 1900 0 1900 0 2812 0 3156 0 3935 0 4472 0 4878 05175 0.5367 0.5472 0.5506 0.5506 0.2683 0.5506

0 0 0 0

Scheduled 5.4

12.1 14.8 19.4 23.2 22.2 21.6

0 0 0 0

Inlet duct.

Burner.

Design Point Performance

N = 16,500 rpm

w/6 - 43 lbm/sec

PR = 7.7 (maximum)

264

AE DC.TR-77-80

A

ACC

ACR

ARR

ARZ

AZ

a

C

C d

CF (r)

C P

C V

CXFC

CXFR

D

DLR

EPR

NOMENCLATURE

Area

Projected area of control volume circumferential

surface in the circumferential direction

Projected area of control volume circumferential

surface in the radial direction

Projected area of control volume radially-facing

surface in the radial direction !

Projected area of control volume radially-facing

surface in the axial direction

Control volume area normal to axial direction

Acoustic velocity; blade spacing; a constant

Blade chord length

Blade drag coefficient

Blade lift coefficient

Stage characteristic correction factor for

radial variation

Specific heat at constant pressure

Specific heat at constant volume

Circumferential crossflow coefficient

Radial crossflow coefficient

Blade drag force

Dynamic lift amplitude ratioed to quasi-steady-

state lift

Stage equivalent pressure ratio, Eq. (47)

265

AE DC-T R-77-80

ETR

e

F

FC

FR

FZ

f

f(...)

g(...)

H

HC

HR

HZ

IMP

IMPR

i

im

ip

J

jm

Stage equivalent temperature ratio, Eq. (46)

Internal energy

Force of compressor blading and cases acting on

fluid, including wall pressure area force

Circumferential force acting on fluid in control

volume

Radial force acting on fluid in control volume

Force of compressor blading and cases acting on

fluid in the axial direction

Frequency

A function of ...

A function of ...

Total enthalpy flux

Total enthalpy flux transported across control

volume circumferentially-facing boundary

Total enthalpy flux transported across control

volume radially-facing boundary

Total enthalpy flux transported across control

volume axial-facing boundary

Impulse function, Eqs. (5) and (160)

Ratio of stage exit to stage entry impulse

functions

Axial location index

Adjacent axial location, i-1

Adjacent axial location, i+l

Radial location index

Adjacent radial location, j-i

266

JP

k

L

M

ME

MFF

MFFE

N

0 ( . . . )

P

PR

PS

Q

R

Re

REI

r

r ~ s

s

T

TR

TS

t

t

U

AE DC-T R-77-80

Adjacent radial location, j+l

Reduced frequency based on full blade chord

length, Fig. ii

Total control volume length; blade lift force

Mach number

Equivalent Mach number

Mass flow function, Eq. (74)

Equivalent stage downstream mass flow function,

Eq. (73) f

Compressor rotor speed

Order of magnitude of ...

Stagnation (total) pressure

Stage total pressure ratio

Static pressure

Rate of heat addition to control volume

Gas constant

Reynolds number

Reynolds number index

Radius; coordinate in the radial direction

Root-mean-square amplitude

Coordinate in the circumferential direction

Stagnation (total) temperature

Stage total temperature ratio

Static temperature

Time

Torque on a fluid element

Axial velocity

267

AE DC-TR-77-80

V

V

vol

W

WB

WC

WR

WS

WZ

W

X

Y

Y

Z

Velocity components other than axial (designated

by subscript)

Radial velocity component in three-dimensional

model development

Volume

Mass flow rate; airflow rate

Compressor bleed flow rate

Mass flux across circumferentially-facing

control volume boundary

Mass flux across radially-facing control volume

boundary

Stage shaft work added to fluid in control

volume

Mass flux across axially-facing control volume

boundary

Circumferential velocity component in three-

dimensional model development

Energy function, Eq. (8)

Time derivative of a volume integral, Eq. (150)

A general dependent variable

Axial coordinate

Greek S~mbols

B

A

Angle of attack

Flow direction angle relative to axial direction

Ratio of specific heats

A difference

268

AE DC-TR-77-80

Ap

6

£

had

e

m

8

P

~P ~T

Pressure fluctuation amplitude

Ratio of compressor entry total pressure to

standard day, sea-level-static pressure

A constant; an error magnitude

Stage adiabatic efficiency

Angular location; ratio of compressor inlet

total temperature to standard day, sea-level-

static temperature

Angular extent of low pressure region in r

circumferential distortion pattern

Blade chord-to-axial direction angle or stagger

angle

Density

Stage pressure coefficient

Stage temperature coefficient (stage loading

parameter)

Angular rotor speed; circular frequency

Superscripts

Average or effective value

P Pertaining to pressure

T Pertaining to temperature

Subscripts

0,1,2,...

ad

avg

crit

Location or station designators

Adiabatic

Average value

Critical value

269

AE DC-T R-77-80

dyn

exit

H

i

im

ip

ijk

J

jm

Jp

k

km

kp

L

l o c

m

m a x

m i n

n

P

radial avg

ref

rel

Dynamic value

Pertaining to control volume exit plane

High

Axial location index

Adjacent axial location, i-I

Adjacent axial location, i÷1

Three-dimensional location index, refers to

value at location i, j, k

Radial location index

Adjacent radial location, j-1

Adjacent radial location, j+l

Circumferential location index

Adjacent circumferential location, k-1

Adjacent circumferential location, k+l

Low

Local value

Modified form; minus one

Maximum value

Minimum value

Last location, e.g., in, last axial location

in control volume

Plus one

Average value along a radius, average at

constant 8

Reference value

Relative value, usually relative to rotor

blade motion

270

AEOC-TR-77-80

ring avg

stall

SS

SW

wh

Z

Average value along a circle, average at

constant r

Value at stall

Steady-state value

Swirl

Wheel

Axial component

271