Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

1988SaliversonPhD PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 212

Saliveros, Efstratios (1988) The aerodynamic performance of the

NACA-4415 aerofoil section at low Reynolds numbers.


MSc(R) thesis.

http://theses.gla.ac.uk/3050/

Copyright and moral rights for this thesis are retained by the author

A copy can be downloaded for personal non-commercial research or


study, without prior permission or charge

This thesis cannot be reproduced or quoted extensively from without first


obtaining permission in writing from the Author

The content must not be changed in any way or sold commercially in any
format or medium without the formal permission of the Author

When referring to this work, full bibliographic details including the


author, title, awarding institution and date of the thesis must be given

Glasgow Theses Service


http://theses.gla.ac.uk/
theses@gla.ac.uk
"THE AERODYNAMIC PERFORMANCE OF

THE NACA-4415 AEROFOIL SECTION AT LOW

REYNOLDS NUMBERS"

A Thesis

Submitted to the Faculty of

Engineering for the Fulfilment of

the Requirements for the Degree

of

Master of Science, (M. Sc. )

by

Efstratios Saliveros, B.Sc.

Department of Aerospace Engineering

University of Glasgow

Glasgow G12 8QQ

November 1988

n© 1988, E. Saliveros"
PREFACE

The work described in this Dissertation was carried


out by the author at the Department of Aerospace
Engineering, University of Glasgow, between October 1986,
and September 1987, and is original in content except
where otherwise stated.

Department of Aerospace Engineering


University of Glasgow
Glasgow G12 8QQ

E. Saliveros

November, 1988

i
ABSTRACT

In this experimental investigation, the performance


and the boundary layer characteristics of the NACA-4415
aerofoil section were examined for an incidence range of

-5.10o~a~22.90o and for the Reynolds number range of


50,OOO~Re~600,OOO. Chordwise static pressure distributions
were obtained, from which aerodynamic force and moment
coefficients, namely CN, CT, and CMc/4, were calculated
using a simple Trapezoidal Rule method. These pressure
distributions proved to be useful for the identification
and location of the various boundary layer phenomena which
occurred around the aerofoil. The "surface oil" flow
visualisation technique was also used and photographs were
obtained to record the various flow states over the upper
surface of the aerofoil. The nominal two-dimensional data
obtained in this study were compared with those from other
facilities and previously tested aerofoils at the
University of Glasgow. These latter comparisons were for
the GU25-5(11)8, GA(W)-l and NACA-0015 aerofoil sections.

ii
ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my


supervisor Dr. R. A. McD. Galbraith for his recommendation
of this project and his valuable advice, assistance and
guidance throughout this work.

I would also like to express my profound gratitude to


the Head of the Department of Aerospace Engineering,
Professor B. E. Richards, without whose support and
encouragement I would not have had the opportunity to work
on this research project.

lowe many special thanks to Mr. G. Kokkodis for his


advice and help during the early experimental stages of
this work, and to Drs. F. N. Coton, A. Kokkalis and to Mr.
R. D. Gordon for their useful suggestions and comments.

I am also thankful to the technicians, Messrs David


J. Perrins, James C. Carr and John A. Kitching, for their
efforts, and the goodwill that they have shown to me,
assisting in any way they possibly could and for providing
me with all the necessary equipment and tools when I really
needed them. Special praises go to Mr. Kitching for his
excellent job of constructing and polishing the model, to
Mr. David M. Whitelaw and to the staff of the Photographic
Department for their efforts of developing the pictures
that appear in this work.

Finally, I would like to thank my parents for their


tremendous support and encouragement that they have given
me throughout my life and to my wife Catherine for the
patience and understanding that she has shown during my
studies at university.

iii
LIST OF CONTENTS

Page
PREFACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
LIST OF CONTENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
NOMENCLATURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

CHAPTER

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

1.1 Background Information ... . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Boundary-Layer Characteristics/Separation Bubble ... 4

1.3 Scope of the Present Study ..... . . . . . . . . . . . . . . . . . . . . 8

II . EXPERIMENTAL APPARATUS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.1 Wind Tunnel Facilities .. . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.2 Wind Tunnel Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.3 Computer Facilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.4 Static Pressure Measurements . . . . . . . . . . . . . . . . . . . . . 13

2.5 Flow Visualisation and Photographic Equipment .... 15

I I I. EXPERIMENTAL PROCEDURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 7

3.1 Static Pressure Distribution Measurements . . . . . . . . 17

3.2 Experimental Limitations . . . . . . . . . . . . ....... ...... 26

3.3 Flow Visualisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

IV. PRESENTATION AND DISCUSSION OF RESULTS . . . . . . . . . . . . . 31

4.1 Performance as a Function of Angle of Attack ..... 32

iv
4.1.1 Introductory Comments . . . . . . . . . . . . . . . . . . . . . . . 32

4.1.2 Re = 50 , 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.1.3 Re = 75, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4.1.4 Re 100 , 00 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.1.5 Re ~ 125, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.2 Pressure Coefficient Variation with Reynolds

numbers at Constant Angles of Attack .... ......... 49

4.3 Flow Visualisation Over the Upper Surface of

the NACA-4415 Aerofoil. . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.4 CN' CMc/4 and CT with Variation of Angles of

Attack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.5 CNmax Variation with Reynolds Numbers ..... ....... 61

4.6 Comparison of the NACA-4415 Aerofoil Section

Characteristics with Existing Data .. ............. 63

4.7 Comparison of the NACA-4415 Aerofoil Section

Characteristics with GU25-5(11)8, NASA GA(W)l

and NACA-0015 Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

V. CONCLUSIONS AND RECOMMENDATIONS FOR FURTHER

STUDIES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

5.1 Summary of Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

5.2 Recommendations for Further Studies . ............. 76

5.3 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

APPENDIX: Manufacture and Assembly of NACA-4415 Aerofoil


Mode 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
REFERENCE S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

v
LIST OF TABLES

TABLE Page

1. Coordinates of NACA-4415 Aerofoil Section ......... 90

2. Locations of Pressure Tappings on NACA-4415


Aerofoil Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

3. Two-Dimensional Angle of Attack Summary . . . . . . . . . . . 92

4. Estimated Locations of Laminar Separation Points on


the Lower Surface of a NACA-4415 Aerofoil Section
Using a Viscid-Inviscid Analysis Method ..... ...... 93

5. Useful Information About Past and Present wind


Tunnels Used to Test NACA-44l5 Aerofoil Sections .. 94

vi
LIST OF FIGURES

Figure Page

1.1 Chord Reynolds Number versus Flight Velocity


for a Variety of Natural and Man-Made Flying
Objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
1.2 A Typical Laminar Separation Bubble Formed over
the Upper Surface Of an Aerofoil .... ............. 95
1.3 Turbulence Effects on Stall Hysteresis (a) and
Pressure Distribution (b) for the Wortmann
FX-63-137 Aerofoil Section. (Re=100,000) ......... 96
1.4 Turbulence Effects on Stall Hysteresis (a) and
Pressure Distribution (b) for the Wortmann
FX-63-137 Aerofoil Section. (Re=200,000) ......... 96
1.5 Acoustic Effects on Stall Hysteresis (a) and
Pressure Distribution (b) for the Wortmann
FX-63-137 Aerofoil Section. (Re=200,000) ......... 96
1.6 Sketch of a Laminar Separation Bubble ............ 97
1.7 Four Different Types of Static Stall ............. 98
1.8 Typical Aerofoil Pressure Distribution with
Laminar Separation Bubbles .... ................... 98
1.9 Experimental Static Pressure Distribution over
the NACA-4415 Aerofoil at an Incidence of -5.10 0
and at a Reynolds Number of 298,051 .............. 99
2.1 A Plan View of the Glasgow University's Low
Speed Wind Tunnel . .............................. 100
2.2 Cross-Sectional View of the Working Section of
the Glasgow University's Low Speed Wind Tunnel .. 101
2.3 Positions of the Pressure Tappings Around the
NACA-4415 Aerofoil Model . . . . . . . . . . . . . . . . . . . . . . . . 102
2.4 Pressure Distribution Comparison Between Two
Wortmann Aerofoils of Having Different Pressure
Tap Configurations at 16 0 Angle of Attack and
Re=200, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

vii
2.5 Pressure Distribution Comparison Between Two
Wortmann Aerofoils of Having Different Pressure
Tap Configurations at 8 0 Angle of Attack and
Re=80,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
3.1· Sequence of Events During Pressure Distribution
Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.2 Schematic of Instrumentation used for the Data
Acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
3.3 Section Lift and Profile Drag Coefficients Versus
Angle of Attack for Rc =150,000 for the Smooth
Lissaman Airfoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
3.4 Section Lift and Profile Drag Coefficients Versus
Angle of Attack for Rc =150,000 for the Smooth
Miley Airfoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
3.5 The NACA-4415 Aerofoil Installed in the Glasgow
University's Low Speed Wind Tunnel .. ....... ..... 108

4.1.1 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=50,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

4.1.2 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=75, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

4.1.3 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=100,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

4.1.4 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=125, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

4.1.5 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=150, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

4.1.6 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=175,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

viii
4.1.7 3-D Plots of Cp vs x/c vs a for the Lower (a) and
Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=200,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

4.1.8 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=250, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

4.1.9 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=300, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

4.1.10 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=35 0 , 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

4.1.11 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=400, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

4.1.12 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=450,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

4.1.13 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=500,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

4.1.14 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=550, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

4.1.15 3-D Plots of Cp vs x/c vs a for the Lower (a) and


Upper (b) Surface of a NACA-4415 Aerofoil Section
and Re=60 0,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.1.16 Pressure Distribution Around the NACA-4415
Aerofoil Section at Incidences Close to Complete
Stall (Re=175, 000) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
4.1.17 Pressure Distribution Around the NACA-4415
Aerofoil Section at Incidences Close to Complete
Stall (Re=300, 000) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

ix
4.1.18 Pressure Distribution Around the NACA-4415
Aerofoil Section at Incidences Close to Complete
Stall (Re=400, 000) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.1.19 Locations of the Various Boundary Layer Phenomena
on the Upper Surface of the NACA-4415 Aerofoil
Section at Re=75,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
4.1.20 Locations of the Various Boundary Layer Phenomena
on the Upper Surface of the NACA-4415 Aerofoil
Section at Re=100, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
4.2.1 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=-5.100 ... 129


4.2.2 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=-3.100 ... 129


4.2.3 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=-1.100 ... 130


4.2.4 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=-0.100 ... 130


4.2.5 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a= 0.90° ... 131


4.2.6 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a= 2.90° ... 131


4.2.7 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a= 4.90° ... 132


4.2.8 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a= 6.90° ... 132


4.2.9 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a= 9.90° ... 133


4.2.10 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=10.900 ... 133

x
4.2.11 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=11.90o ... 134


4.2.12 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=12.90 o ... 134


4.2.13 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=13.90 o ... 135


4.2.14 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=14.90o ... 135


4.2.15 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=15.90o ... 136


4.2.16 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=16.90o ... 136


4.2.17 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=17.90o ... 137


4.2.18 3-D Plot of Cp vs x/c vs Re for the Upper, Surface

of the NACA-4415 Aerofoil Section at a=18.90o ... 137


4.2.19 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=19.90o ... 138


4.2.20 3-D Plot of Cp vs x/c vs Re for the Upper Surface

of the NACA-4415 Aerofoil Section at a=20.90o ... 138


4.2.21 Typical Surface Pressure Distributions With and
without Separation Bubble . . . . . . . . . . . . . . . . . . . . . . . 139
4.3.1 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at a=-5.10o ....... 140


4.3.2 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at a=-1.10o ....... 141

xi
4.3.3 Flow Visualisation Photographs of the Upper
Surface of the NACA-44l5 Aerofoil Section at

Various Reynolds Numbers and at U= 2.90 0 ....... 142


4.3.4 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at U= 6.90 0 ....... 143


4.3.5 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

various Reynolds Numbers and at U= 9.90 0 ....... 144


4.3.6 A sequence of Flow Visualisation Photographs of
the Upper Surface of the NACA-4415 Aerofoil at

Re=600,000 and u=9. 90 0 .......................... 145


4.3.7 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

various Reynolds Numbers and at U=11.90 0 ....... 146


4.3.8 A sequence of Flow Visualisation Photographs of
the Upper Surface of the NACA-4415 Aerofoil at

Re=600,000 and at U=11.90 0 .....•••••••.•....... 147


4.3.9 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at U=13.90 0 ....... 148


4.3.10 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at U=15.90 0 ....... 149


4.3.11 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at U=17.90 0 ....... 150


4.3.12 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at U=18.90 0 ....... 151


4.3.13 Flow Visualisation Photographs of the Upper
Surface of the NACA-4415 Aerofoil Section at

Various Reynolds Numbers and at U=19.90 0 ....... 152


xii
4.3.14 A sequence of Flow Visualisation Photographs of
the Upper Surface of the NACA-4415 Aerofoil at

Re=450,OOO and at a=19.90 o . . . . . . . . . . . . . . . . . . . . . 153


4.3.15 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=150,000 . . . . . . . . . . . . . . . 154
4.3.16 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=200,000 . . . . . . . . . . . . . . . 155
4.3.17 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=250,OOO . . . . . . . . . . . . . . . 156
4.3.18 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=300,000 . . . . . . . . . . . . . . . 157
4.3.19 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=350,000 . . . . . . . . . . . . . . . 158
4.3.20 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=400,000 . . . . . . . . . . . . . . . 159
4.3.21 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=500,000 . . . . . . . . . . . . . . . 160
4.3.22 Comparison between Cp and Flow Visualisation Data
Regarding the Locations of the Various Boundary
Layer Phenomena occurring on the Upper Surface of
the NACA-4415 Model at Re=600,000 . . . . . . . . . . . . . . . 161

xiii
4.4.1 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=50,000 . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4.4.2 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=75,000 . . . . . . . . . . . . . . . . . . . . . . . . . . 163
4.4.3 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=100,000 . . . . . . . . . . . . . . . . . . . . . . . . . 164
4.4.4 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=125, 000 . . . . . . . . . . . . . . . . . . . . . . . . . 165
4.4.5 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=150,000 . . . . . . . . . . . . . . . . . . . . . . . . . 166
4.4.6 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=175,000 . . . . . . . . . . . . . . . . . . . . . . . . . 167
4.4.7 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=200,000 . . . . . . . . . . . . . . . . . . . . . . . . . 168
4.4.8 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=250,000 . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.4.9 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=300, 000 . . . . . . . . . . . . . . . . . . . . . . . . . 170
4.4.10 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=350,000 . . . . . . . . . . . . . . . . . . . . . . . . . 171
4.4.11 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=400,000 . . . . . . . . . . . . . . . . . . . . . . . . . 172
4.4.12 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=450,000 . . . . . . . . . . . . . . . . . . . . . . . . . 173

xiv
4.4.13 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient variation With Angle
of Attack at Re=500,000 . . . . . . . . . . . . . . . . . . . . . . . . . 174
4.4.14 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=550, 000 . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.4.15 Normal Force, Quarter Chord Pitching Moment and
Tangential Force Coefficient Variation With Angle
of Attack at Re=600, 000 . . . . . . . . . . . . . . . . . . . . . . . . . 176
4.4.16 Quarter Chord Pitching Moment Coefficient versus
Angle of Attack for the Wortmann FX-63-137
Aerofoil (R c =100,000) . . . . . . . . . . . . . . . . . . . . . . . . . . 177
4.4.17 NACA-23012 Aerofoil Data for AR=4.0 . . . . . . . . . ... 177
4.4.18 Lift Coefficient versus Uncorrected Incidence of
a NACA-23012 Aerofoil at a Reynolds Number of
350,000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
4.4.19 Measured Aerodynamic Characteristics of a NACA-
23012 Ai r f 0 i 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
4.4.20 Comparison of Results for the Occurrence of the
'kink' on CN and CMc/4 Curves with those Defining
Fully Laminar Flow Using Coton's Method . . . . . . . . . 179
4.4.21 Two Dimensional Lift and Drag Coefficients
versus Angle of Attack, Re=80,OOO .............. . 179a
4.5.1 Maximum Normal Force Coefficient Variation with
Reynolds number . . . . . . . . . . . . . . . . . . . . . . . . . · ... ···· 179
4.5.2 Lift Coefficient Variation with Incidence and
Reynolds Number of a NACA-4415 Aerofoil Tested
at stuttgart (1962-72) . . . . . . . . . . . . . . . . . . . . . . . . . 180
4.5.3 Lift Coefficient Variation with Incidence and
Reynolds Number of a NACA-4415 Aerofoil Tested
at NACA VDT (1934) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
4.5.4 Lift Coefficient Variation with Incidence and
Reynolds Number of a NACA-4415 Aerofoil Tested
at NACA LTT (1945) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
4.5.5 Normal Force Coefficient Variation with Incidence
and Reynolds Number of a NACA-4415 Aerofoil Tested
at University of Glasgow (1987) . . . . . . . . . . . . . ... 183

xv
4.5.6 Lift Coefficient Variation With Incidence of a
Wortmann FX-63-137 Aerofoil Tested at a Reynolds
Number of 200, 000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
4.6 Comparison of Lift and Quarter Chord Pitching
Moment Coefficients of the NACA-4415 Aerofoil
Obtained at Different Test Environments and Wind
Tunnels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
4.7.1 Normal Force Coefficient Contours for the NACA-
4415 Aerofoil with Incidence and Reynolds
Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
4.7.2 Normal Force Coefficient Contours for the GU25-
5(11)8 Aerofoil with Incidence and Reynolds
Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
4.7.3 Normal Force Coefficient Contours for the NASA
GA(W) -1 Aerofoil with Incidence and Reynolds
Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
4.7.4 Normal Force Coefficient Contours for the NACA-
0015 Aerofoil with Incidence and Reynolds
Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
A1. Construction of the NACA-4415 Model . ........... . 191

xvi
NOMENCLATURE

c Aerofoil chord length, rom


CL Lift coefficient
CMc/4 Quarter chord pitching moment
CN Normal force coefficient
CNmax Maximum normal force coefficient
Cp Pressure coefficient
CT Tangential force coefficient

dCMc/4/da Quarter chord pitching moment curve slope

dCN/da Normal force curve slope


g Gravitational acceleration, m/s2
H Atmospheric pressure height, rom Hg
h21 Differential pressure height between atmospheric
and static pressure, rom H20
Differential pressure height between settling
chamber and working section, rom H20
Wind tunnel calibration constant
Static pressure based at surface of aerofoil, Nm 2
Atmospheric pressure, Nm 2
Pressure at wind tunnel settling chamber, Nm 2
Pressure at wind tunnel working section, Nm 2
Free stream dynamic pressure, Nm 2
Specific gas constant, kJ/KgK
Reynolds number based on aerofoil chord
Tunnel air speed, °c
Free stream velocity, m/s
x Chordwise co-ordinate

xvii
Greek Symbols

IX Angle of attack, deg.

IXg Geometric angle of attack, deg.

£B Total blockage factor

P Density of air, Kg/m 3

Pm Density of mercury relative to water

Pw Density of water, Kg/m 3

Subscripts

a air
c chord
g geometric
m mercury
s settling chamber
w working section, water
CHAPTER I

INTRODUCTION

1.1 BACKGROUND INFORMATION

In aerodynamic applications, the low Reynolds number

regime is usually taken as that for which the chord

Reynolds number falls below 1,000,000. In the past, study

of this flow regime has not been actively investigated

since most aerodynamic vehicles operate at much higher

Reynolds numbers. The last twenty years, however, have seen

a growing interest in low Reynolds number flows since a

number of applications have evolved requiring aerofoil

sections that operate at low Reynolds numbers. A working

knowledge of the associated flow phenomena is therefore

necessary.

Low Reynolds number applications occur when any

combination of the following conditions is present:

a) low free stream velocity

b) low air density

c) small aerofoil chord.

These conditions are found to exist on remotely piloted

vehicles (RPV's) used for surveillance, sampling and

- 1 -
monitoring in both military and scientific roles,

operating at high altitudes, mini-RPV's flying at low

altitudes, high altitudes jet-engine fan blades, etc.

Additional applications are found in the inboard sections

of helicopter, propeller, and in wind turbine rotors.

Figure 1.1 succinctly illustrates the various Reynolds

number regimes. Efforts in designing low Reynolds number

aerofoil sections which possess high aerodynamic

eff iciencies have been effective together with much

experimental work to determine the performance of existing

aerofoils at those Reynolds numbers (Ref.l).

Flow behaviour at chord Reynolds number less than

1,000,000 is widely known to involve some significantly

different characteristics when compared to higher Reynolds

number flows. These flow characteristics create

difficulties relating to the management of the aerofoil

boundary layer as well as difficulties associated with

accurate wind tunnel measurements (Refs 2,3). Lissaman

states, that as a general criterion there exists a critical

Reynolds number of about 70,000 below which aerofoil

performance is very poor and above which impressive

improvements are observed (Ref.4). Very important areas of

concern are the occurrence and behaviour of the leading

edge laminar separation bubbles and the associated

phenomenon of transition from laminar to turbulent flow

(F igure 1.2). It is well known that separat ion is highly

sensitive at low Reynolds numbers and plays a very

- 2 -
important role in determining the development of the
boundary layer which, in turn, affects the overall

performance of the aerofoil.

On close examination of the wind tunnel test data

below the 1,000,000 range of Reynolds numbers using models

with the same aerofoil sections, it is possible to observe

that there is a number of inconsistencies among the test

results. This can be attributed to a variety of causes

including inaccurate measurement techniques, or due to

solid and wake blockage effects, and differences between

test environments. These blockage effects are discussed in

detail in References 5 and 6. In the past, researchers

have often been puzzled by other researchers questioning

test data accuracy, data acquisition techniques, data

measurement reliability, model accuracy or even tunnel

corrections. Mueller et al (Ref.7) describes some of the

methods used to identify the level of free stream

disturbances and their influence on the performance of low

Reynolds number aerofoils.

The problems associated with obtaining accurate wind

tunnel data for aerofoil sections at low Reynolds numbers

are compounded by the extreme sensitivity of the boundary

layers to the free stream disturbance environment. The

disturbance environment present in the test section of a

low speed wind tunnel is usually determined by the free

stream turbulence levels, acoustic phenomena and mechanical

-3-
vibrations. Although these disturbances may be reduced and,

to some extent, controlled, they cannot be eliminated. If

proper care is taken to reduce or account for such

phenomena, then more meaningful results may be obtained. By

taking account of turbulence intensity levels and acoustic

disturbances Sumantran et al (Ref.8) obtained considerable

differences in the test data and in particular the stall

hysteresis loop and pressure distributions (Figures 1.3,

1.4 and 1.5). These Figures illustrate that the range of

turbulence levels between 0.02% and 0.2% is quite important

whilst above a value of 0.2% may have negligible effects.

At these increased values maximum lift coefficients appear

to remain unaffected. Similar results were also obtained by

Mueller (Ref.9), during an investigation on the performance

of two aerofoils influenced by free stream disturbances at

low Reynolds numbers.

1.2 BOUNDARY LAYER CHARACTERISTICS I SEPARATION BUBBLE

The phenomena of boundary layer separation and

transition at low Reynolds numbers have been known to be

very important for a long time because of their effect on

the aerofoils performance. With the existence of turbulent

flow over an aerofoil, drag is increased and so a desirable

design for an aerofoil, therefore, would be to maintain

laminar flow over a large proportion of its length. Such a

design would lead to higher lift and lower drag values. At

low Reynolds numbers, however, where the flow tends to

- 4 -
remain laminar it may have been expected to produce flows

tending towards this aim. Unfortunately, the inability of

the laminar boundary layer to sustain large adverse

pressure gradients leads to the separation of the flow from

the aerofoil surface within the pressure recovery region.

This results in large losses of lift and increased drag.

If the laminar shear layer separates from the surface

and undergoes transition to turbulent flow, and the

turbulent shear layer has sufficient entrainment to

reattach, then the well known "separation bubble" will be

formed. The separation bubble is defined as that region in

which slowly recirculating air is trapped between the

separation and re-attachment points. A simple diagram of

the separation bubble is shown in Figure 1.6. The length of

the separation bubble is defined as the distance between

the separation and re-attachment points and is usually

expressed in non-dimensional form as a percentage of the

chord length.

Laminar separation bubbles have been studied for many

years (Refs 10-15). Generally their behaviour is used to

describe the performance and stalling characteristics of

aerofoils. These stalling characteristics can be classified

into four different categories:

a) Trailing edge stall

b) Leading edge or short-bubble stall

c) Long bubble or thin aerofoil stall

-5-
d) Combination of both trailing edge and leading edge
stall.

The characteristics of these categories are illustrated in


Figure 1.7.

"Trailing edge stall" generally occurs on

moderately thick aerofoils (~10% x/c). It is identified by

forward movement of the turbulent separation point as

incidence is increased causing a decrease in the lift

curve slope, and a gradual loss of lift beyond the maximum

without causing a sharp drop in the lift coefficient.

"Leading edge" or "short bubble stall" occurs when a short

bubble situated just downstream of the leading edge suction

peak "bursts", causing gross flow separation. This occurs

when the leading edge adverse pressure gradient is too

severe for the flow to re-attach. The third type, the "long

bubble" or "thin aerofoil stall", is described as the

movement of the bubble's re-attachment point towards the

trailing edge of the aero foil with increasing angle of

attack. The long bubble increases in length gradually until

it covers the entire aerofoil. The maximum lift is

relatively small compared with other types of stall and if

the angle of attack is increased further this will lead to

lift reduction. The final type of stall occurs when the

aero foil experiences a combinat ion of trail ing edge

separation and leading edge short bubble bursting. At

angles of attack just prior to the leading edge bubble

bursting, the lift curve slope dips as the lift decreases

- 6-
only gradually due to significant trailing edge separation.

With increases in angle of attack the leading edge bubble

bursts resulting in a sudden decrease in lift and a large

increase in drag. In Reference 11, Chappell calls such

behaviour a "combined stall" due to the occurrence of both

forward movement of the turbulent separation point and the

leading edge bubble bursting.

Separation bubbles are typically described as

"long" or "short" depending on their relative lengths. A

"long" bubble covers a separated region of about 20%-40% of

the chord, while the length of a "short" bubble covers only

a few percent. A short bubble generally makes little impact

on the pressure distribution around an aerofoil (Figure

1.8). Lissaman (Ref.4) also uses the above criterion in

defining long bubbles, but states that the short bubble

could not form unless the chord Reynolds number is greater

than 100,000.

Many of the separation bubbles identified in this

study occurred as far downstream as 65% or even 75% of the

chord with turbulent re-attachment very close to the

trailing edge. These bubbles were "long" according to their

length criterion (20%-40% of the chord), but became shorter

in length and moved towards the leading edge at increased

incidence. When the aerofoil was set at negative angles of

attack (e.g. a=-5.1 0 , Re=300,OOO) two separation bubbles

were formed, a "long" and a "short" one. Figure 1.9,

- 7-
illustrates a pressure distribution with the two separation

bubbles formed on the upper and lower surfaces clearly

shown. The long bubble formed near the trailing edge on the

upper surface and the short one close to the leading edge

on the lower surface of the aerofoil.

During the stalling process, on the NACA-4415

aerofoil, a combination of both trailing edge separation

and leading edge short bubble bursting was observed,

indicating that the aerofoil's stalling characteristic fell

into the fourth category as mentioned earlier. The

two-dimensional behaviour of the NACA-4415 aerofoil with

increasing angle of attack and Reynolds number is described

in Chapter IV.

1.3 SCOPE OF THE PRESENT STUDY

The investigation carried out in this study was to

identify possible advantageous characteristics of the

NACA-4415 aerofoil section at Reynolds numbers below or

equal to 600,000. This work is also part of the continuing

research programme at the University of Glasgow dedicated

to revealing the performance and boundary layer characteri-

stics of aerofoils which are currently being used on the

rotor blades of various wind turbines and helicopters. In

addition to this investigation, comparison was made with

three other aerofoils previously tested using the same wind

tunnel facilities. It is hoped that what was observed will

- 8-
be valuable to future researchers and may give an

indication of the phenomena which other aerofoil sections

operating in the same low Reynolds number regime, would

experience.

Static pressure measurements were taken to obtain

values of normal force and pressure drag coefficients as

well as pitching moment coefficients. An assessment of the

first order boundary layer characteristics was made and it

revealed that the most important feature was the separation

bubbles developed on the upper and lower surfaces of the

model. Flow visualisation was also carried out to examine

the flow phenomena over the aerofoil's upper surface and

useful information was recorded in the form of photographs

which were subsequently used in the analysis of the test

data.

- 9-
CHAPTER II

EXPERIMENTAL APPARATUS

All experimental data presented in this report were

obtained using the facilities of the Aerospace Engineering

Department's Laboratory at the University of Glasgow. The

test model, a NACA-4415 aerofoil section was used and its

performance determined over a wide range of low chord

Reynolds numbers. Static pressure distributions were

obtained using pressure tappings which were connected via

three electronic selector boxes, to a micromanometer and

thence to a DEC MINC (PDP 11/23) mini-computer system.

Aerodynamic force and moments coefficients were obtained by

integrating the static pressure distribution of the

aerofoil. The "surface oil" flow visualisation technique

was also used and photographs were obtained to record the

various flow states that occurred around the aerofoil. The

technique provided useful information for a better

understanding of the aerofoil's performance and boundary

layer characteristics.

2.1 WIND TUNNEL FACILITIES

All tests were conducted in the Aerospace Engineering

-10-
Department's medium sized low speed wind tunnel which is an

atmospheric-pressure closed-return type and has a

rectangular cross section of 1.143m width and 0.838m height

(Figures 2.1 and 2.2). Its upper and lower walls contain

graduated turntables for mounting the model vertically.

This was done in such a way as to provide rotation about

the quarter chord axis.

Since the tunnel operated at atmospheric pressures


only, the Reynolds number was varied by means of changing

the tunnel airspeed. The minimum and maximum air flow

velocities obtained in the test section were approximately

2.5 and 30.0 m/s respectively. Kokkalis, (Ref.16), investi-

gated the turbulence intensity level and found that in the

longitudinal direction it was 0.4% while in the lateral

direction it was 0.6%. Both of these components were

measured at the centre of the working section and under a

free stream velocity of 10.0 m/s.

2.2 WIND TUNNEL MODEL

The model used in these experiments had a NACA-4415


aerofoil section. The construction of this model was

carried out at the Aerospace Engineering Department's

manufacturing and assembly facilities during the period

December 1986 to February 1987. Full details of the

construction of this model are given in the Appendix. The

chord length was 300.0 mm, and the span 838.0 mm giving an

-11-
aspect ratio of 2.8. The dimensions were chosen so they
would match those of GU25-5 (11) 8, NASA GA (W) -1 and

NACA-0015 models previously tested under the same

conditions in the same wind tunnel facility (Refs 17 and

18). The aerofoil's co-ordinates are listed in Table 1.

The NACA-4415 aerofoil section was chosen because it

belongs to the NACA-44XX family of aerofoil sections which

are widely used for the rotor blades of horizontal axes

wind turbines . Galbraith et al (Ref. 19), provides a

tabulated list not only of the NACA-44XX aerofoil series

but also of the NASA GA(W) -1 and NACA-0015 aerofoil

sections showing their wide applicability to various types

of wind turbines.

2.3 COMPUTER FACILITIES

In the present study, data acquisition and data

reduction for all pressure measurements was accomplished

with the aid of a DEC MINC (PDP 11/23) mini-computer. The

data acquisition system consisted of PDP 11/23 processor

interfaced with two RX02M diskette drives. Each diskette

has a capacity of 512 kB. One of the diskettes was used as

the system's device (DYO:) containing the necessary

operating system software while the other (DY1:) was used


for data and programs storage. The console used with this

system, while the computer operated in the data acquisition

mode, was a DEC VT105. An IBM computer system and an EPSON

-12-
MX-80 printer were also connected with the computer system

for the production of draft pressure distribution plots.

Analog to digital conversion was accomplished by a

12-bit A/D converter (MNCAD). In conjunction with the A/D

converter a programmable clock (MNCKW) was also used. Both

the A/D converter and the clock were plugged into the

MINC's chassis in a configured mode.

All the collected and reduced data files were

transferred from the diskettes to a VAX 11/750 computer

main-frame via a lengthy process. The VAX 11/750 coupled

with a VERSATEC plotter and using GINO graphics routines,

produced high quality plots. The VAX 11/750 computer system

used to produce the required plots and permanent storage of

the data for Data Base Management (D.B.M).

2.4 STATIC PRESSURE MEASUREMENTS

Aerofoil pressure distributions were obtained using

a specially constructed model with pressure tappings

mounted on its mid-span position on the upper and lower

surfaces along the chord length. Pressure data were

obtained for ranges of 50,000 to 600,000 chord Reynolds

numbers and -5.10 0 to 22.90 0 of angles of attack.

A total of sixty pressure tappings were placed

around the aero foil so a good assessment of the chordwise

-13-
pressure distribution could be achieved. Thirty nine of

those were placed on the upper surface with the remaining

twenty one on the lower surface of the model. The

locations of these tappings were measured using a vernier

height gauge and are listed in Table 2. Figure 2.3

illustrates the shape of the aerofoil's cross section

together with the positions of the pressure tappings on

both the upper and lower surfaces. The pressure tappings

were staggered over the first 11.4% and 8% of the chord

from the leading edge on the upper and lower surfaces

respectively, and at 22% from the trailing edge on the

upper surface. Staggering the pressure tappings was

deliberate to avoid any interference which might exist

between a downstream pressure tapping and an upstream one

(Ref.1). Figures 2.4 and 2.5 show clearly wide differences

in pressure distribution measurements between an "in-line"

and "staggered" pressure tapping model tested under the

same wind tunnel conditions. As seen in Figure 2.4, the

"in-line" pressure tap model has a lower suction peak and

fails to indicate the existing leading edge separation

bubble. Figure 2.5 also demonstrates that the pressure

tappings, placed "in-line", cause earlier transition.

The use of two electronic micromanometers was

required to measure free stream dynamic pressures and

differential pressures at each pressure tapping. Both

micromanometers are of MDC-FC002 and FC012 types and have

ranges of +/-19.99 and +/-199.9 rnrn of H 2 0.

-14-
The accuracy 0 f the MDC- FC 002 mi cromanometer,

calibrated by the manufacturers using precision water

column manometers, is +/-1%, whilst for the FC012 type it

is 0.2% or 0.3% depending on range of pressure. Their

linearity is +/-0.5% or +/-1.0%, and their output voltage

signal is 0-2 VDC or 0-5 VDC respectively (see Reference


22) .

Alongside the micromanometers, three selector boxes


were provided to accommodate all pressure tubes. Each

selector box has a maximum of twenty pressure fittings

attached at its rear panel together with an output

pressure port connected to the input of the FCO 12

micromanometer. These selector boxes enabled an automatic

selection of each pressure tapping so that the

corresponding pressure measurement could be carried out.

This automatic selection of the tappings was achieved by

using an IEEE Standard 488-1975 Bus Controller mounted in

the PDP 11/23 computer. The communication path between the

PDP 11/23 and the selector boxes was provided by the IEEE

Bus interface cable.

2.5 FLOW VISUALISATION AND PHOTOGRAPHIC EOUIPMENT

In addition to the pressure coefficient and force

measurements on the NACA-4415 model tested much useful

information about the boundary layer was obtained by

-15-
visually observing the nature of the fluid flow past the

surface of the model. For this purpose, the surface oil

flow visualisation technique was considered. This technique

was chosen because of its successful application on

different aerofoil models tested in the same wind tunnel by

previous researchers (Refs 17,18,20 and 21).

This flow visualisation technique was performed using

Odina-oil, Saturn Yellow "Dayglo" fluorescent powder and

liquid paraffin for a thinner. The viscosity of the mixture

was adjusted by trial and error until a suitable ratio was

obtained. This mixture was then applied on the upper

surface of the model by careful stippling using an ordinary

sponge. Extra care was taken so the oil mixture was

uniformly distributed on the whole area of the upper

surface of the model. Once the solution was applied to the

model it was illuminated using two ultra-violet light

sources. They made the visualisation solution appear bright

yellow, and therefore enabled the user to observe and study

the flow patterns as they developed.

A NIKON FE-2 50mm camera was used to obtain still

photographs of the fluorescing powder once the pattern had

developed. The camera was equipped with ultra-violet and

polarising filters so that only the visible light from the

powder impinges on the photographic emulsion. Photographs

were obtained using ILFORD XPl-400 ASA film.

-16-
CHAPTER III

EXPERIMENTAL PROCEDURE

3.1 STATIC PRESSURE DISTRIBUTION MEASUREMENTS

The chordwise pressure distributions were obtained at

the mid-span of the specially constructed pressure model

using the PDP 11/23 data acquisition system, two electronic

micromanometers and three selector boxes. An ordinary

thermometer and a barometer were also used to measure the

wind tunnel air temperature and atmospheric pressure

respectively during each test. Two programs were used

extensively in this study for the collection and reduction

of the data, namely AEROFL.BAS and CNTM41.FOR. The data

collection program (AEROFL.BAS) written in BASIC computer

language and the data reduction program (CNTM41. FOR)

written in FORTRAN language. The main reason of writing

AEROFL.BAS in BASIC was that the three selector boxes could

only operate under BASIC language commands. These programs

are a modification of the programs GEOR.BAS and CNDM.FOR

used by Kokkodis (Ref.lS). The modification was necessary

to improve time consumption and accuracy of the reduced

data. A set of programs was also written to present the

reduced data files in a graphical form. All the plotting

-17-
programs were written in FORTRAN and, together with GINO

graphics routines, a presentation of chordwise pressure

distributions in two and three dimensional form, as well as

force and moments coefficients versus angles of attack, was

available.

The main assumption made for this part of the study

was that the airflow remained uniform, steady and

incompressible in its entire journey through the tunnel.

Using the above assumption the pressure coefficient is

defined as

(3.1)

where P1 refers to static pressure on the surface of the

aerofoil

P2 refers to atmospheric pressure and

q refers to free stream dynamic pressure.

The pressure difference (P1 -P2) can also be written as

(Ref.23)

(3.2)

and the free stream dynamic pressure as

-18-
(1/2) 'Pa.v2 (3.3)

where k is referred to as the wind tunnel calibration

factor. For the wind tunnel concerned, k equals to 1.18.

Thus, equation (3.1) can be re-written as

(3.4)

Therefore to determine the pressure coefficient it was only

necessary to measure two pressure differences, both

measured in mm H20.

The flow chart illustrating the sequence of events

during the chordwise pressure distribution measurements is

shown in Figure 3.1. A schematic representation of the

various electronic instruments used and their inter-

connections is also shown in Figure 3.2. The free stream

dynamic pressure was measured in terms of the difference

between a total head and a static pressure reading using

the MDC-FC002 electronic micromanometer. The static

pressure was measured by means of an orifice in the wall of

the working section, well upstream of the model. The total

head pressure was measured through another orifice placed

in the tunnel's settling chamber wall upstream of the

contraction. The two orifices were connected to the

measuring head of the micromanometer via rubber tubes and

-19-
the pressure difference was then recorded. In addition, a

second micromanometer (FC012) measured the pressure

difference between static pressures from each tapping on

the model and the atmospheric pressure.

Before each set of tests, a warm up time of twenty to

thirty minutes was allowed so that the electronics of the

two micromanometers as well as the selector boxes and the

AID converter were brought up to desired operational

temperatures. At the beginning and end of each test, the

wind tunnel air temperature and the atmospheric pressure

were recorded. The air temperature was measured using an

ordinary thermometer inserted through an orifice in the

side wall of the tunnel upstream of the test section. The

atmospheric pressure was obtained using a mercury

barometer. The barometer height ((H) in mm Hg) and air

temperature ((Ta) in °C) were substituted into expression

(3.5) which determined the wind tunnel air density

(Ref. 24) .

Pair (3.5)

The determination of wind tunnel air density was necessary

in order to reduce to the minimum any experimental errors

concerning the air velocity and hence the free stream

dynamic pressure and Reynolds number.

As mentioned earlier in Chapter II, three selector

-20-
boxes were used to accommodate all pressure tubes. Each

selector box had twenty pressure ports at its rear panel.

The connection of the pressure tubes was made in such a way

that tube labelled as number one in selector box I

corresponded to the pressure tapping nearest to the

trailing edge on the upper surface of the model. The

pressure tapping nearest to the trailing edge on the lower

surface corresponded to tube number sixty connected to

selector box III. An IEEE Bus cable provided the

communication link between the PDP 11/2.3 computer and the

selector boxes. The boxes were connected to the Bus by

multiple conductor cables. With the help of the IEEE Bus

cable the selection of a particular pressure port, and

therefore of a pressure tapping on the model, could be

carried out automatically by the acquisition program.

Each selector box also has a main output pressure

port which allows the pressure of a selected pressure

tapping to be transmitted into the FC012 micromanometer.

The manometer could then measure the pressure difference

between the atmospheric pressure and the static pressure of

the selected pressure tapping. The reading was then fed

into the A/D converter of the PDP 11/23 computer as an

analog voltage signal. The selector boxes were entirely

governed by the computer which allowed complete computer

control of the experiment for the duration of each test.

A "test" consisted of the chordwise static pressure

-21-
distribution data taken at mid-span of the model for one

Reynolds number condition and for one angle of attack.

During each test, the data acquisition program stepped

through the model once, sampling each pressure tapping

forty times in a period of one second. The sampling process

for each test began at the trailing edge and proceeded

along the upper surface to the leading edge then on to the

lower surface pressure tappings starting from the leading

edge and progressing to the trailing edge. The cycle of the

selection of the forty samples for each pressure tapping

was carried out twice and two average values were obtained.

If the difference between the two averages was between

+/-2% of the first average then this value was stored on

the system's diskette before progressing to the next

tapping. However, in case the above convergence criterion


was not satisfied, then the whole process of taking new

values would be repeated up to a maximum of twenty times.

If still no convergence was obtained after the twenty

cycles then the last average value was recorded together

with a warning that this particular pressure tapping did

not converge. Therefore, it was easy to observe that the

time taken for each test varied in the present

investigation. During some tests the time used was

considerable. This occurred due to the large flow

fluctuations for very low Reynolds numbers (see Figures

4.1.1 and 4.1.2) .

The selected pressure measurements from each tapping

-22-
were stored on diskettes so that the reduction process of
the data could be carried out after the completion of the

tests. Before any experiments were begun, several runs were

carried out to check for possible pressure tube leaks at

various wind tunnel speeds and angles of attack as well as

the data acquisition process. After the completion of each

test the corresponding pressure distribution was displayed

on the screen of the DEC VT105 monitor. This allowed a

checking of any pressure abnormalities that might have


occurred during the experiment and therefore a re-run, for
that particular angle of attack and Reynolds number, was

carried out.

During each set of tests, the wind tunnel operated


at the desired speed until the completion of the tests.

This had an effect on the dynamic pressure which with

increasing angle of attack was reduced due to wind tunnel

blockage effects. The dynamic pressure was originally set

with zero angle of attack for each set of tests. Generally,

the dynamic pressure decreased for each set of tests up to

a maximum of 3% of its initial value, prior to complete

stall.

In general, a complete set of tests for the

NACA-4415 aero foil at one Reynolds number condition

included angles of attack prior to the commencement of

positive lift to angles beyond full stall. Measurements

were taken for at least every degree of incidence. The

-23-
Reynolds number range varied from 50,000 to 600,000 with a

step increment of 25,000 for the first 200,000. After

200,000 Reynolds number, the step increment increased to

50,000 until it reached 600,000 .. A grand total of 597 tests

were performed for 15 different Reynolds number conditions.

The CNTM41. FOR computer program was used for the

integration of pressure distributions to obtain normal

force and pressure drag coefficients, together with leading

edge and quarter chord pitching moment coefficients. The

integration of pressure distributions was based on a

trapezoidal rule approximation method. When using chordwise

pressure distributions this method is relatively accurate

provided there is close spacing of pressure tappings around

the aerofoil. The program also stored pressure coefficients

data in files formatted for plotting using the available

software in the Aerospace Engineering Department's VAX

11/750 computer library.

Throughout this investigation, the generated force

and moment coefficients remained uncorrected for two-

dimensional blockage effects and streamline curvature.

There were two main reasons for leaving these coefficients

uncorrected. Firstly, because no attempt was made to

measure the wake behind the aerofoil by using a wake

traverse method, and secondly, because of the limited

validity of the derived two-dimensional equations, applied

for such corrections. These equations are only valid for

-24-
flows over aerofoils which are wholly subsonic and fully

attached (Refs 5 and 6) .

Such wake measurements would have been essential for

obtaining a value of the total blockage factor, £s' at

different geometric angles of attack. Although the flow

over the present aerofoil matched the subsonic criterion,

it did not match the fully attached flow. This was because

of the presence of a separation bubble on either the upper

or lower surfaces and trailing edge separation at

incidences greater than 80 - 90 . Full details of the

expressions involved for correcting wind tunnel aerodynamic

force and moment coefficients data are given in References

5 and 6.

The only correction made in this study, however, was

of the angle of attack. Kelling (Ref.21) discovered in his

investigation that the flow approaching the test section is

yawed by approximately +/-0.6 degrees depending on the

direction in which the test incidence is measured. Here,

the positioning of the model in the working section

contributed to a negative flow yawing angle. Therefore, the

actual angle of attack was obtained by subtracting 0.6

degrees from the measured geometric incidence, i.e.

ex (3.6)

-25-
3.2 EXPERIMENTAL LIMITATIONS

The main experimental limitation experienced in this

study was the inability to obtain data for the NACA-4415

aerofoil at angles of attack greater than 19.40° and for

550,000 Reynolds number. This problem was encountered due

to the fact that the FC012 electronic micromanometer went

out of range. Pressure values exceeding the micromano-

meter sI operational limit were usually observed on the

upper surface and very close to the leading edge. When this

occurred no further collection of data was allowed and the

test was terminated.

Exactly the same problem occurred for the 600,000

Reynolds number and for angles of attack greater than

13.40°. For a matter of interest however the micromanometer

was set to manual with a maximum off-set value of -40.7 mm

H20 as the zero setting. This allowed the micromanometer to

extend its operational limit to +240.6 and -159.2 mm H20.

Once again, however, it proved impossible to reach full

stalling angles. By pure coincidence, the maximum angle of

attack obtained for both the 550,000 and 600,000 Reynolds

number cases was 19.40 0 • One might speculate that the data

collected for Reynolds number equal to 600,000 and for

angle of attack greater than 13.40 0 is not accurate enough

and should be considered with skepticism. Figure 4.4.15

however, illustrates that there are no major discrepancies

-26-
of the aerodynamic force coefficients against angle of

attack plots for angles of attack greater than U=13.40o and

therefore the data could be treated as such of lower

Reynolds numbers.

Generally, in an experimental investigation of the

aerodynamic performance of an aerofoil, a consideration

must be made for every possible aspect which might affect

its overall performance. Such an aspect is the hysteresis

effect occurring near the stalling angles of attack. The

aerodynamic forces are strongly dependent on this

hysteresis phenomenon and on the direction at which the

stalling angle is approached. Hysteresis is of practical

importance because it could strongly affect the recovery

from stall and/or flight conditions.

Two possible hysteresis loops exist, the "high-lift"

or "clockwise" and "low-lift" or "counterclockwise"

hysteresis. By increasing the angle of attack the lift and

drag forces are increased. When stall is finally reached

the lift experiences a large drop while drag experiences a

great increase. Reducing the angle of attack slightly the

former values of the aerodynamic forces are not restored.

Instead, the angle of attack may have to be reduced by

several degrees before lift and drag revert to values

obtained under conditions of increasing angle of attack.

This is known as the "high-lift" or "clockwise" hysteresis

-27-
loop and mainly happens to aerofoils experiencing early
transition caused by a separation bubble.

For the "low-lift" or "counterclockwise" hysteresis

loop, the lift and drag experience sudden increases at a

certain angle of attack. In this case, reducing the angle


of attack, lift and drag keep increasing until maximum

values are obtained. Reducing it even further causes lift

and drag forces to revert to values obtained with

increasing angles of attack. This type of hysteresis loop

is mainly caused by the increase of the long bubble and its

abrupt collapse to a short bubble. Figures 3.3 and 3.4 show

that the Lissaman aerofoil exhibits the "high-lift"

hysteresis loop while the Miley produces the "low-lift"

hysteresis loop.

In this study, however, considering that there is a


very high turbulence intensity level (0.5%) in the wind

tunnel used and following the observations of Sumantran et

al and Mueller (Refs 8 and 9), it is thought that

hysteresis effects may be considered to be negligible.

Therefore, no attempt was made to show any hysteresis

effects which might occur on the lift curve slopes.

3.3 FLOW VISUALISATION

Flow visualisation was accomplished using the surface

oil-film technique. This technique was carried out on the

-28-
NACA-4415 aerofoil model with all the pressure orifices

sealed. No quantitative pressure measurements were made at

this stage, since the objective was to observe any flow

phenomena which may occur. Photographs were taken of the

model's upper surface at different angles of attack and


Reynolds numbers. This method had been used very

successfully in the past (References 18 and 20) at tunnel

speeds over 10mls but with very limited success at lower

velocities. The interpretation of flow visualisation

photographs taken in the study, as well as in previous

studies for speeds lower than 1 Oml s, indicate that the

surface oil-film flow visualisation technique as used, was

performing at its limit and the oil-film probably altered

the boundary layer characteristics of the aerofoil giving


inaccurate results.

It was found from initial photographic studies that

good results could be obtained for all ranges of flow

conditions if the camera shutter speed was 1/15 second and

the aperture was set at f:2.0. All photographs in this

study were obtained using these camera settings. Film


developing and printing was accomplished using the

Aeronautics Department photographic facilities as well as

those of the Photographic Department.

The model was mounted vertically in the test section

(Figure 3.5) and the NIKON FE-2 camera mounted on a tripod

positioned approximately 60cm from the test section side

-29-
wall. Typically, it was necessary to shut down the tunnel

between tests in order to apply the correct oil film. Extra

care was taken to ensure that this oil film was uniformly

distributed over the aerofoil. Without delay the angle of

attack was set and the tunnel speed was brought up from

zero to its desired value. The oil started to move along

the chord in streaks. Since the model was mounted vertical-

ly in the test section, gravitational effects gave a down-

ward bias to the flow pattern. The flow pattern was allowed

to develop until no further flow changes were observed. The

surface flow patterns were then photographed.

Leading edge separation bubbles were clearly visible

and noted as narrow vertical bands of trapped re-

circulating oil covering a small area of the model surface

along the span (Figure 4.3.1~. Long bubbles were

characterised by a separation line followed by a region

where the oil remained stationary. The location of the

re-attachment point was identified as the line immediately

behind the bubble, indicated by the beginning of a wider

dark band representing fully attached flow (Figure 4.3.2).

Following the dark band, and for angles of attack greater

than 11.90 0 , an uneven line developed which represented the

turbulent separation front.

The main advantage of this flow visualisation method

is that the different flow fields formed on the surface of

the aerofoil could easily be distinguished.

-30-
CHAPTER IV

PRESENTATION AND DISCUSSION OF RESULTS

This chapter presents a selection and detailed

discussion of the most pertinent results obtained from both

the chordwise static pressure distribution and flow

visualisation measurements on a NACA-4415 aerofoil. From

these measurements, the two-dimensional performance of the

aerofoil with respect to angle of attack, a, and chord

Reynolds number, Re c ' was determined.

Due to the large amount of data collected during the

present investigation, pseudo-three-dimensional represent-

ations were developed to illustrate the static pressure

distributions over the aerofoil. They are plotted against

the full range of angles of attack considered at constant

Reynolds numbers and vice versa. Aerodynamic force and

moment coefficients, used for studying the performance of

an aerofoil, were obtained by integrating the chordwise

pressure distribution using a simple trapezoidal rule

method. These coefficients are presented as a function of

angle of attack for each Reynolds number tested. No

corrections, to the results throughout this study for the

-31-
effects of two-dimensional wake, solid blockage, and of

streamline curvature (refer to Section 3.1), were applied

to the data.

Further, a surface flow visualisation technique was

used to assist understanding of the flow mechanisms which

might have affected the behaviour of the boundary layer

over the upper surface of the aerofoil.

The analysis of the results obtained from the

pressure and flow visualisation measurements helped to

clarify the two dimensional performance of this aerofoil in

the Reynolds number range considered.

4.1 PERFORMANCE AS A FUNCTION OF ANGLE OF ATTACK AND

REYNOLDS NUMBER

4.1.1 Introductory Comments

The chordwise static pressure distribution plots for

the NACA-4415 aerofoil at a Reynolds number range of 50,000

to 600,000 and for angles of attack ranging from -5.10 0 to

22.90 0 are illustrated in figures 4.1.1 to 4.1.15 and 4.2.1

to 4.2.20. The first set shows the pressure variations over

the upper and lower surfaces at various incidences while

the Reynolds number is kept constant. The second set is a

presentation of the pressure differences that occur on the

upper surface of the aerofoil for the Reynolds number range

-32-
examined, at constant angles of attack. Both sets of

figures offer a good indication of the presence of laminar

separation bubbles, as well as, the locations of turbulent

boundary layer separations, on the upper or lower surface

of the model. The locations of laminar separation and

turbulent re-attachment points, however, could not be

always ascertained from these figures. It was found that

the performance of the aerofoil at these low Reynolds


numbers was dictated by the positions of laminar

separation, transition and turbulent re-attachment. Turbu-

lent boundary layer separation, however, had an effect upon

the generation of the aerodynamic forces and pitching

moments. Due to the observed repeatability of the boundary

layer phenomena over the test range, only the first three

Reynolds number cases (i.e. 50,000, 75,000 and 100,000) are

examined in detail.

For Reynolds numbers as low as 50,000 and 75,000, the

static pressure distribution plots show that the pressure

along the chord length fluctuates quite strongly for all

angles of attack tested. These pressure fluctuations are

possibly due to environmental disturbances in the wind

tunnel. Previous studies (Refs 7,8) have shown that for low

Reynolds number regimes, environmental disturbances have

played a significant part in the behaviour of the boundary

layer and, consequently, the performance of an aerofoil.

Such fluctuations in pressure make the study of the

boundary layer itself a very formidable task. The amount of

-33-
data collected from the present tests, however, allows a

detailed discussion of the flow phenomena around the

aerofoil to be carried out with some confidence.

Identification of the various boundary layer

phenomena from the pressure distribution plots was carried

out using the following criteria. Areas of relatively

constant pressure indicated the position of laminar or

turbulent flow separation. The location of flow transition

and that of turbulent re-attachment were assessed from the

classical interpretation of the pressure profiles

exhibiting bubble characteristics. Arena et al (Ref.12)

describes turbulence re-attachment as that point where the

pressure, when it is fully recovered, is nearly equal to

the value measured for a turbulent boundary layer over an

aerofoil with no presence of a separation bubble (Figure

4.2.21) .

4 .1.2 ~c 50,000

The pressure distributions over the upper and lower

surface of the aerofoil for a Reynolds number of 50,000 are

shown in Figures 4.1.1(a) and (b) respectively. Examining

these figures it may be observed, that on the lower

surface, a laminar separation bubble is formed close to the

leading edge. At -5.10 0 , the length of the bubble appears

to be between 23% and 33% of the chord, which according to

the length criteria, described in Section 1.2, the bubble

-34-
is considered to be "long". With increasing incidence,

however, the separation bubble persists, increases in

length slightly and moves aft towards the trailing edge. As

indicated by the pressure distribution plots, the bubble

seems to disappear at a higher incidence, but due to

pressure fluctuations, it is very difficult to assess the

incidence at which this occurs.

In contrast to the lower surface, the behaviour of

the boundary layer over the upper surface (Fig.4.1.1(b))

appears to be more simple. Due to the pressure fluctuations

little can be said about the behaviour of the boundary

layer, except that the suction peak moves closer to the

leading edge with increasing incidence and that stalling of

the aerofoil is believed to occur at an incidence of 6.90 0 .

The location of stall is taken from the pressure

distribution plots which show that aft of x/c=0.15 the

pressure coefficient is nearly constant, indicating that

leading edge flow separation has occurred. with increasing

incidence the aerofoil remained stalled and laminar

separation progresses forward from x/c=0.15 at u=6.90 0 to

approximately x/c=0.03 at u=22.90 0 .

From the aerodynamic force and moment coefficient


plots, shown in Figure 4.4.1, the normal force coefficient,

eN' appears to vary in a linear manner over the range from

-5.10 0 to 0.90 0 of angles of attack. Over the same range,

-35-
however, the quarter chord pitching moment coefficient,

CMc/4' decreases in magnitude with increasing incidence.

Zero normal force is obtained at an angle of attack of

about -1.50 0 . As the angle of attack is increased the CN

curve slope gradually decreases. From the incidence where

stall occurs (6.90 0 ), CN rises very slowly. Due to

fluctuations of the CN curve between 6.90 0 and 22.90 0 ,

however, only an average value of dCN/du=1.71 has been

estimated.

Between angles of attack of 0.90 0 and 13.40 0 , CMc/4

coefficient maintains an approximately constant value of

-0.09. As incidence increases to 22.90 0 the magnitude of

CMc/4 decreases; exceeding -0.13 at high angles of attack.

Minimum tangential force coefficient of 0.037 occurs at

U=-o .10 0 . From U=4. 90 0 to 22.90 0 , however, it remains

relatively constant, having an average value of 0.015

(Fig.4.4.1)

The aerofoil's performance at this Reynolds number

can be characterised as very poor, especially for angles of


attack greater than 6.90 0 where it approximates that of a

flat plate.

-36-
4.1. 3 Reo 75,000

At this Reynolds number the performance is dominated

by the formation of separation bubbles on both the upper

and lower surfaces. Pressure fluctuations persist over the

test range, but are much less than those observed at

Re=50,000. Upper and lower surface static pressure

distributions and aerodynamic coefficients are presented in

Figures 4.1.2 and 4.4.2 respectively.

For a=-5.10 0 , a leading edge separation bubble is

formed on the lower surface (Fig. 4.1.2(a)). It appears to

be shorter than that observed for Re=50,000 and occupies

about 18% of the chord. Additionally, laminar separation

occurs at x/c=O. 08 while transition and turbulent flow

re-attachment are approximately at x/ c=O. 23 and x/ c=O. 28

respectively. Also, turbulent flow separation appears to be

present at the x/c=O. 88. For increased incidences, the

bubble continued to exist but with slight increase in

length until around a=-0.10 0 where it has disappeared.

The behaviour of the flow over the upper surface

(Fig.4.1.2(b)) is dominated by the formation of a

persistent separation bubble over almost the whole range of

angles of attack tested up to the stall. For small angles

of attack as low as -5.10 0 , laminar separation is at about

x/c=0.56 with transition possibly occurring close to the

-37-
trailing edge. At increased incidence, the laminar

separation point moves forward and the flow does not

re-attach until u=-2. 10 0 . There is then an abrupt

re-attachment close to the trailing edge forming an upper

surface long separation bubble. This bubble occupies

approximately 50% of the chord with laminar separation

occurring at about x/c=O. 50 and transition at x/c=O. 84.

With further increases in incidence, the separation bubble

migrates forward and shortens to 40% at u=3.90 0 and 35% at

U=9.90 0 . Just before stall, however, its overall length is

around 13% chord. Turbulent flow separation first appears

around the 12.40 0 incidence and moves upstream towards the

leading edge with increasing angle of attack. At U=16.40 0

turbulent flow separation has moved to x/c=0.62 causing a

38% chord trailing edge separation. Leading edge

separation, caused by the "burst" of the bubble occurred at

U=16.90 o . The bubble's progression, discussed above, is

illustrated in figure 4.1.19 where the location of laminar

separation, transition, turbulent re-attachment and

turbulent separation are shown.

The formation of the upper and lower surface bubbles

as well as their disappearance played a significantly

important part into the performance of the aerofoil. Prior

to the upper surface bubble formation, the dCN/dU curve

-38-
slope increased gradually to a large value of approximately

9.50 (Fig.4.4.2). When the bubble formed, at a=-2.10 0

(Fig.4.1.2(b)), the magnitude of the slope reduced to 6.7

(a decrease of nearly 30%) and remained so until there was

no lower surface bubble at a=-0.10 0 . The disappearance of

the lower surface bubble probably caused the dCN/da value

to decrease even further by nearly 13% to an approximate

value of 5.85. Over the same variation of incidence, the

quarter chord pitching moment coefficient decreased in

magnitude reaching a minimum of -0.106 at a=0.900. In the

range -1.10 0 <a<4.900, however, dCMc/4/dafluctuates between

-0.105 and -0.097. The tangential force coefficient,

however, is different and increases in magnitude to a

maximum of 0.03 at a=-3.100 and then decreases gradually

with increasing incidence. From a=4.900 the magnitude of

the dCN/da slope decreases reaching a minimum at 12.40°, at

which CN attains its maximum value of approximately 1.41.

Over the same incidence range (8.90° to 16.40°) the

quarter chord pitching moment coefficient increases in a

somehow unsteady manner with a sharp rise of about 22%

between 11.90° and 12.90°. This increase is thought to have

occurred due to the initiation of trailing edge flow

-39-
separation. Similarly, CT continues to decrease further

with increasing incidence, reaching a minimum value of

-0.274 at U=15.90 o .

Leading edge flow separation (see figure 4.1.2(b))

causes the aerofoil to stall at an angle of attack of

U=16.90 o . This is manifested by a large decrease in both

the normal force and pitching moment coefficients, together

with a sharp rise in the tangential force coefficient (Fig.

4.4.2). At this incidence, CN and CMc/4 have dropped to

0.84 and -0.114, while CT reached a value of 0.013.

At this chord Reynolds number, the performance of

the aerofoil has improved considerably compared to the

50,000 case.

4.1.4 Reo =100,000

Figure 4.1.3 illustrates the static pressure

distributions around the aerofoil at various angles of

attack. For this Reynolds number and the remaining cases

the pressure fluctuations which dominated the first two

Reynolds numbers cases, are greatly reduced and the

identification of boundary layer phenomena was made easier.

From the pressure distributions it may be observed that

separation bubbles are formed on both the upper and lower

-40-
surfaces. From an incidence of -5.10 0 to -0.10 0 , the

aerofoil has a lower surface separation bubble close to the

leading edge. This then extends downstream with increasing

incidence (Fig. 4.1.3(a)). At u=-0.10 0 the bubble has moved

to mid-chord occupying approximately 30% of the chord

length and laminar separation has moved to x/c=0.48, with

transition and turbulent re-attachment at about x/c=O. 68

and x/c=0.78 respectively. For U=0.90 0 , however, the

corresponding pressure distribution shows that the bubble

has disappeared and this had a significant effect on the

aerofoil's performance which exhibited a noticeable "kink"

in the normal force and pitching moment curves. Detail

discussion of the "kink" is given in Section 4.4.

The behaviour of the upper surface boundary layer is

again influenced by the existence of a separation bubble

Fig.4.1.3(b). Static pressure distribution plots, however,

show that at u=-5.10 0 laminar separation and transition of

the flow appear to occur at approximately x/c=0.62 and 0.92

respectively, with no indication of turbulent re-

attachment. As the incidence is increased, the flow tends

to re-attach and at an angle of -3.10 0 is located at

x/c=O. 94. This forms a long separation bubble covering

about 40% of the chord. At this incidence, laminar

separation and transition have moved to their new positions

of x/c=0.56 and 0.86 respectively. The bubble then exhibits

similar trends, with increasing incidence, to that of the

-41-
Re=75,000 case.

It is observed that it forms at a lower incidence and

monotonically migrates towards the leading edge, whilst

decreasing in length as incidence is increased. Trailing

edge separation also takes place at a=12. 90 0 and moves

rapidly upstream with increasing incidence. However, just

prior to stall (a=16.40 0 ), the bubble develops close to the

leading edge covering 10% of the chord, while the trailing

edge separation point has progressed to x/c=O. 56. The

migration of the bubble along the chord together with that

of the trailing edge separation location is illustrated in

figure 4.1.20.

By examining the aerodynamic coefficients (Fig.

4.4.3), it is noticeable that the dCN/da value changed in

magnitude at four different incidences before stall was

eventually reached. In the first stage, from -5.10 0 to

-2.10 0 , its value was found to be about 7.3. At a=-2.100

the flow on the upper surface re-attached forming a long

bubble. It could be argued that the change of flow

behaviour on that surface caused the reduction of the curve

slope to a magnitude of 5.9, a decrease of nearly 20%. The

curve slope remained constant in magnitude until the

leading edge bubble on the lower surface disappeared at

-0.10 0 .

-42-
The re-attachment of the flow also produced distinct

effects on dCMc /4/ da curve slope. Prior to flow

re-attachment (-2.10 0 ) the CMc/4 curve has a negative

slope, whilst for greater incidence values the slope became

positive. It remained positive until CMc/4 reached a

maximum value of -0.030 at 14.40 0 except, when the lower

surface bubble disappeared, a slight "kink" is obvious on

both the CN and CMc/4 curves. This "kink" is more distinct

in greater Reynolds numbers (Figs 4.4.4 to 4.4.15) and is

discussed in Section 4.4.

For angles of attack between 0.90 0 and 3.90 0 , dCN/da

(Fig.4.4.3) obtained a value of 5.44. The flow in this

incidence range is probably fully attached, apart, of

course, for the bubble formation. A further incidence

increase to 7.90 0 increased the rate of movement of

transition and re-attachment points over their predecessors

(Fig.4.1.3(b)). This caused the boundary layer to become

turbulent prematurely, which probably affected the CN curve

slope by reducing it to 4.6. Further increases in incidence

lead to additional losses in deN/da. A maximum normal force

coefficient of 1.375 was achieved at an angle of attack of

12.400. For greater incidences the CN was reduced as the

trailing edge separation increased until when leading edge

-43-
separation (bubble burst) occurred at 16.65 0 , CN and CMc/4

dropped from 1.294 to 0.829 and -0.034 to -0.110

respectively, while CT increased from -0.260 to 0.005.

4.1.5 Ego ~125,000

Static pressure distributions over the aerofoil for

Reynolds numbers between 125,000 and 600,000 are shown in

Figures 4.1.4 to 4.1.15. Examining these figures, it may be

observed that the trends are similar to those discussed in

the last two Reynolds number cases. An attempt to discuss


each figure individually would have provided no extra

information about the behaviour of the boundary layer, than

has already been discussed. Therefore, only a brief

discussion of the most obvious changes in pressure will be

described here. As far as the aerodynamic parameters are

concerned they appear to undergo some noticeable changes,

particularly in moderate angles of attack (less than 4 0 )

and beyond CNmax. A discussion of their behaviour will be

presented in Section 4.4.

Although the most unusual discontinuities in

pressure distribution occurred on the upper surface of the

aerofoil, some slight changes were also observed on the

lower surface. There the pressure distributions (Figs

4.1.4(a) to 4.1.15(a)) indicate that, at incidences below

zero degrees, a separation bubble forms close to the

-44-
leading edge. With increasing Reynolds number, however, it
l

shortens by approximately 8-12% per 50,000 of Reynolds

number. When the angle of attack is gradually increased the

bubble travels rapidly downstream until it disappears at a

small incidence. The disappearance of this bubble is


observed to take place at an incidence not higher than
1 . 90 0 for the lower Reynolds number range (i. e. ,

125,000~Rec~200,000) and not later than 0.90 0 for the

higher regimes. Further increases in angle of attack, did

not feature any unusual discontinuities in Cp until the

incidence at which stall occurred. It is understood, that


when this happened, the suction peak on the upper surface
collapsed and the centre of pressure shifted aft towards
the trailing edge. This is clearly noticeable by the sudden
forward movement of the stagnation point towards the

leading edge.

As was described above, a laminar separation bubble


was also formed on the upper surface (Figs 4.1.4(b) to

4.1.15(b)). It was observed that its formation was near the


trailing edge and, as the angle of attack increased, it
migrated upstream. When formed, it is described as a "long"
bubble, according to bubble length criteria. With
increasing incidence, however, it travels upstream and

becomes a "short" bubble covering only a small percentage


of the chord. This is clearly illustrated in figure

4.1.10 (b) where, for example, at a Reynolds number of


350,000 the bubble is about 25% of the chord at an

-45-
incidence of -5.10 0 and is reduced to approximately 4% at

17.90 0 . Apart from these changes in location and size of

the laminar separation bubble, turbulent flow separation

also takes place near the trailing edge at a certain angle

of attack; depending on the Reynolds number.

This separation moves towards the leading edge with

increasing incidence at a faster rate than that of the

laminar separation, transition and re-attachment points. It

is apparent, that when the separation bubble establishes

itself within the first 15% of the chord, its rate of

displacement slows substantially from about 10% to less


than 4% of the chord per degree of increasing incidence,

depending on the Reynolds number (Figs 4.1.4 (b) to

4.1.15(b)). From these figures it is also noticed that the

higher the Reynolds number, the earlier trailing edge

separation occurs.

In the present investigation, it became apparent


that the boundary layer around the aerofoil proved to be

very sensitive at these low Reynolds number regimes and

particularly for incidences close to stall. Occasionally,


during a single static pressure measurement test, the

aerofoil experienced two different flow conditions, i.e. a


fully attached and a fully separated. The occurrence of
these flow phenomena is clearly shown in figure 4.1.16(b).

This Figure illustrates that at a Reynolds number of

approximately 300,000 and at an angle of attack of 18.90 0 ,

-46-
a short bubble is situated close to the leading edge on the
upper surface of the aerofoil. Laminar separation and

transition occur at the locations of x/c=0.035 and 0.055

respectively, while re-attachment, indicated by the end of

the sharp pressure recovery region, occurs at approximately

x/c=0.075. As the pressure continues to increase, past the

re-attachment point, turbulent flow separation takes place

at about x/c=O. 44. After this the pressure coefficient

experiences a sudden decrease in value at approximately


x/c=0.56 and remains relatively constant for the rest of
the chord length.

In contrast with the upper surface, the lower

surface pressure distribution shows that as the pressure

decreases aft of the stagnation point at x/ c=O. 025 , it


experiences again a sudden increase. This occurs around

x/c=0.06 and the pressure decreases steadily from that

point to the trailing edge.

Since the pressure distribution around the aerofoil

was displayed on the computer's visual display terminal

(VT105), immediately after the completion of each test run,

it was possible to check those pressure abnormalities that

had occurred during the experiment before any data analysis

was made. By retesting the NACA-4415 aero foil for the same
incidence (18.90 0 ) and Reynolds number (300,000), it was

observed that the aerofoil was fully stalled (Fig.

4.1.16(c)).

-47-
Comparing the data of figure 4.1.16(b) with that of

the stalled case (Fig.4.1.16(c)), it may be noticed that

the upper surface pressure coefficient values from 0.62 x/c

to the trailing edge, together with those from the leading

edge to 0.05 x/c on the lower surface, are almost

identical. Bearing in mind that the order in which the

static pressure measurements were made (refer to Section

3.1), the most obvious explanation is that the aerofoil was

originally stalled and a flow perturbation triggered the

boundary layer to re-attach. A short time later, the flow

re-separated and subsequently re-attached. From these

figures it is clear that the boundary layer is so sensitive

that it is possible to change its state from fully

separated to fully attached and vice-versa (Figs 4.1.17 and

4.1.18). Such boundary layer behaviour was only observed at

incidences close to stall and can be attributed to the wind

tunnel's environmental disturbances, such as mechanical

vibrations, noise and high turbulence intensity level.

Similar observations were also made by Kokkodis

(Ref.18) for a NASA GA(W)-l and a NACA-0015 aerofoil

sections tested under the same conditions, emphasising the

extreme sensitivity of the boundary layer in these low

Reynolds numbers (0.5x105~Rec~6.0x105) regardless of the

aerofoil section.

-48-
4.2 PRESSURE COEFFICIENT VARIATION WITH REYNOLDS NUMBERS

AT CONSTANT ANGLES OF ATTACK

The pressure distribution over the upper surface of

the NACA-4415 aerofoil for constant incidences and a

variation of Reynolds numbers is illustrated in figures

4.2.1 to 4.2.20. From these figures it may be observed that

the laminar separation point over the upper surface appears

to occur further aft as the angle of attack and Reynolds

numbers increase up to -0.10 0 and 200,000 respectively

(Figs 4.2.1 to 4.2.4). For the same range of angles of

attack and for higher Reynolds numbers, its location moves

upstream, in the way that a conventional laminar separation

point should have behaved.

Further increases in incidence up to 9.90 0 (Figs

4.2.5 to 4.2.9) showed that its position was approximately

within 5% of the chord for all Reynolds numbers considered.

Meanwhile, since transition of the flow occurs earlier with

increasing Reynolds number, turbulent flow re-attachment

also occurs earlier, resulting in a reduced bubble length.

For example at an incidence of -0.10 0 (Fig. 4.2.4), the

bubble occupies approximately 30% of the chord at a

Reynolds number of 125,000. In contrast to this and at a

Reynolds number of 600,000, the bubble covers less than 15%

of the chord. Similar reductions, of more than half of its

original size, were also noticed for the remaining cases

considered. The transition and re-attachment locations,

-49-
which, as described above, occur earlier with increasing

Reynolds number, have similar locations for incidences

between 0.90 0 and 10.90 0 and Reynolds numbers higher than

450,000 (Figs 4.2.5 to 4.2.10). This clearly indicates

that, for these incidences and Reynolds number regimes, the

gross behaviour of the boundary layer is similar. It is

also in this test range that incipient trailing edge

separation occurs (Fig.4.2.9).

For angles of attack of 11.90 0 and 12.90 0 (Figs

4.2.11 and 4.2.12), however, there is anomalous pressure

data, in that the bubble is difficult to distinguish and,

indeed, could be reasonably argued, not to exist. Test

cases on either side of these data and the evidence of flow

visualisation clearly indicate bubble existence (Figs

4 .2.10 and 4.2.13, 4.3. 7 and 4.3. 9) .

For higher angles of attack (a>13.9 0 ), the separation

bubble continues to decrease in length and moves closer to

the leading edge until full stall occurs. At which the

bubble "bursts" resulting in a mixed type of stall since

significant trailing edge separation develops from lower

incidences. Stall of the aerofoil starts initially from the

lower Reynolds number regimes (Fig.4.2.16) and progresses

towards the higher ones until the aerofoil is fully stalled

for all the Reynolds numbers test range (Fig.4.2.20).

Bastedo (Ref.33) also found, during his experimental

-50-
investigation of a Wortmann FX-63-137 aerofoil, similar

difficulties of locating the upper surface separation

bubble from pressure data within a particular incidence


range. Using a direct injection smoke-flow-visualisation

technique illuminated by a laser sheet, however, he was

able to observe that the bubble not only existed but was

also very thin (less than 1mm in thickness) as it moved

forward with increasing incidence. He explains, that as the

bubble becomes thinner it allows transmission of the free

stream pressure, of the external potential flow field, to

the aero foil surface. Therefore, the pressure plateau

normally characteristic of a separation bubble is masked by

this effect. Such an explanation might also be accepted for

the pressure peculiarities observed in the present investi-

gation mentioned above.

4.3 FLOW VISUALISATION OVER THE UPPER SURfACE OF THE


NACA-4415 AEROFOIL

The particular method used to visualise the upper


surface flow fields, was the oil flow method. This

technique has been described in detail previously in

Section 2.5. It has limited applicability, however, (refer

to Section 3.3), and no attempt to use it below a Reynolds


number of 150,000 was made. Therefore, tests were only
carried out at Reynolds numbers between 150,000 and

600,000, for discrete angles of attack. During these tests,

a large number of photographs were collected, from which

-51-
only a representative few are presented here (Figs 4.3.1 to

4.3.14). Full presentation of these pictures is given in

Reference 25. From these photographs, the main features of

the boundary layer, such as laminar or turbulent flow

separation and flow re-attachment, are easily obtained.

It should noted that for these flow visualisation

tests, the tunnel speed was increased from zero to its

desired value after setting the angle of attack. Therefore,

the flow pattern does not necessarily represent the

situation which might be obtained when stall is approached

at constant Reynolds number, by increasing the angle of

attack, as it was done for the pressure measurement tests.

Figure 4.3.1 shows the flow developments that occur

over the upper surface of the aerofoil at various Reynolds

numbers but with a fixed incidence of -5.10 0 . Here the flow

may be termed nominally two-dimensional, except near the

upper model/tunnel interface. The~e minor three-dimensiona-


lity is observed to occur at almost all the Reynolds

numbers considered. The laminar separation bubble can

easily be distinguished, forming just prior to the

mid-chord position and re-attaching near the trailing edge.


Its position across the span is more clearly seen at the
Reynolds numbers of 300,000 and 600,000, with laminar
separation occurring at a much later location compared to

the other Reynolds numbers cases (Figs 4.3.1(b) and (e)).

-52-
At the lower half of the model for the 600,000 case,

as well as, the mid-span position for the 500,000 Reynolds

number (Fig.4.3.1(d)), turbulent oil streaks are observed.

These cut through the separation bubbles and are mainly

caused by lumps of "Dayglo" powder pigments located just

ahead of the bubble which induce transition. Uniform

trailing edge separation is also noticed to have taken

place at approximately x/c=0.98 for a Reynolds number of

300,000 (Fig.4.3.1(b)). As the Reynolds number increases

the length of the bubble is reduced. Such reductions are


caused by the occurrence of earlier re-attachment.

As the angle of attack is increased to -1.10 0

(Fig.4.3.2), the bubble appears to be positioned close to

the mid-chord for all the Reynolds numbers tested.

Comparing the separation bubbles in this figure with those


obtained in the previous angle of attack case, it may be

noticed that they not only become shorter with increasing

Reynolds numbers, but also move upstream with increasing

incidence. The length of the bubble reduces by almost 30%

between 250,000 and 600,000 Reynolds numbers. Trailing edge

separation is also observed to have occurred at about

x/c=0.94 at the Reynolds numbers of 250,000 and 350,000

(Fig.4.3.2 (a) and (b)).

Further increases in incidence up to 9.90 0 (Figs

4.3.3 to 4.3.5), indicate that although the nature of the

flow over the model remains two-dimensional, there is

-53-
little flow separation at the upper model/tunnel wall
junction. The centre of the bubble is located at

approximately 20-30% of the chord, depending on Reynolds

number. Trailing edge separation is present at all Reynolds

numbers tested, and increases with increasing angle of


attack.

At an incidence of 6.90 0 and a Reynolds number of

500,000 (Fig.4.3.4(d)), the bubble in the lower part of the

model is broken due to the presence of a powder particle,

located at x/c=0.03, which trip the boundary layer from

laminar to turbulent. It is also noticed that as the

Reynolds number becomes higher than 400,000 and 300,000 for

angles of attack of 6.90 0 and 9.90 0 respectively, oil

starts to escape from the bubble into the turbulent region

of the flow. This is probably caused by the thinning of the

separation bubble, as described in the previous section. It

is assumed that as the bubble gets thinner with increasing

incidence, the thickness of the oil, which is trapped

inside the bubble, becomes thicker than the bubble itself

and therefore, the shear stresses just above the bubble


drive the oil into the turbulent flow region. This

phenomenon can be seen quite clearly in Figure 4.3.6 where

a sequence of photographs was taken at intervals of

approximately fifteen seconds.

The flow behaviour over the aero foil at an incidence

of 11.90 0 and for various Reynolds numbers, is shown in

-54-
Figure 4.3.7. At this incidence, the laminar separation

bubble is close to the leading edge and its length is

reduced considerably by about 50% when compared to the

corresponding value at 9.90 0 . Turbulent flow separation is

now in the mid-span region around 70-80% of the chord. The


separation line, however, indicates three-dimensional flow

behaviour. Finally, as in the previous case, oil

accumulation in the laminar separation bubble is

significant and some escapes into the general flow

(Fig.4.3.8) .

At angles of attack of 13.90 0 and 15.90 0 (see figures

4.3.9 and 4.3.10), the separation bubble is clearly in the

leading edge region and covers only 4-5% of the chord.

Trailing edge separation is greater than 50% of the chord

and strong vortical flows are evident of the wall/model

interfaces. It may also observed that, between 400,000 and

500,000 Reynolds numbers, flow behaviour differs between

13.90 0 and 15.90 0 incidence. This difference is that, at

13.90 0 incidence, obvious three-dimensional flow exists


over the lower half of the model. At 15.90 0 , however, the
flow is closer to its original nominally two-dimensional

status, and the two vortices end appear to be of similar

extent.

For angles of attack 17.90 0 and 18.90 0 (Figs 4.3.11


and 4.3.12), the flow pattern is similar to that observed

at 15.90 0 , but the separation bubble is almost at the

-55-
leading edge, and the flow reversals covers almost 70% of

the chord. Nominal two-dimensionality persists; even at

lower Reynolds numbers than before (Fig.4.3.11(b)).

The flow development for an incidence of 19.90 0 is

shown in Figure 4.3.13. For Reynolds numbers up to 400,000

(Figs 4.3.13(a), (b) and (c)) the aerofoil is stalled and

the flow separates from the 1% chord location. At higher

Reynolds numbers, however, the flow remains attached for

25% of the chord with the separation bubble still present.

Assymetry of the junction vortices is noticeable; the lower

vortex forming away from the model/tunnel wall junction.

For a Reynolds number of 400,000 (Fig.4.3.14), the

flow, although separated from the aerofoil's surface (Figs

4.3.14(a) and (b)), suddenly changed its status from fully

separated to fully attached (Figs 4.3.14(c) and (d)). A

similar behaviour was also observed during the pressure

measurements. This again illustrates the sensitivity of the

boundary layer at incidences near full stall.

Surface oil flow visualisation technique provided an

interesting correlation with the pressure measurements.

That is, it confirmed, for Reynolds numbers higher than

200,000 the indications given by the pressure distributions

for the existence and behaviour of the laminar separation

bubbles. It may be noticed, however, that laminar

separation, as well as, turbulent re-attachment and

-56-
trailing edge separation, occur earlier with the oil flow

measurements than those deduced from the pressure profiles

(Figs. 4.3.15 to 22) .

The origins of this mismatch are at present unknown

and it is anticipated that in order to resolve the

conflicting data, especially at low Reynolds numbers,

further detailed investigations will be requested.

Such an investigation should include a more

sophisticated analysis of the pressure coefficient data and

a less intrusive method of flow visualisation (perhaps

liquid crystals). Oil flow visualisation is normally highly

informative but at the very low Reynolds numbers

considered, where the boundary layer is very sensitive, the


accumulated oil droplets may have had a severe effect. With

this in mind, the results presented here for Reynolds

numbers less than 600,000 must therefore be viewed with

reservations until the matter is resolved.

4.4 ~NL-CMc/4 AND CT WITH VARIATION OF ANGLES OF ATTACK

The normal force, tangential force and quarter chord

pitching moment coefficients plotted against angle of

attack, are shown in Figures 4.4.1 to 4.4.15. In general,

the trend of the curves is that CN and CMc/4 increase with

-57-
increasing incidence, while CT decreases. This behaviour

continues until the magnitude of the curve slopes reaches a

minimum. For angles of attack prior to the obvious "kink"

at low incidences, however, the dCN/da curve slope is at

its greatest. As the maximum value of CN is approached, the

slope decreases gradually until maximum CN is achieved.

This decrease in slope was caused by both the thickening of

the boundary layer and, primarily, the occurrence of

trailing edge separation (see figures 4.2.1 to 4.2.20).

Increases in incidence above CNmax lead to substantial

decreases in CN. As above, this may be attributed to the

large regions of trailing edge separation on the upper

surface of the aerofoil. Continued increases of incidence

result in sudden loss of CN due to the laminar separation

bubble bursting. Therefore, due to the occurrence of

trailing edge separation and the bursting of the leading

edge bubble, the stall behaviour of the NACA-4415 aerofoil

shows that it is of the combined category; discussed in

Section 1.2.

At Reynolds numbers greater than 200,000 and

incidence values greater than aCNmax (Figs 4.4.5-4.4.13),

the CN curve levels off as the effects of turbulent

separation become manifest. Also, from the force and moment

curves, it may be noticed that the zero normal force

-58-
incidence shifts towards the design value of -4 0 with

increased Reynolds number; this shift is clearly noticeable

between 50,000 and 125,000 Reynolds numbers (Figs 4.4.1 to

4.4.5). The stalling angle may also observed to increase

(Table 3) .

Similar behaviour of the curve slopes is observed at

higher Reynolds numbers, except between 400,000 and

500,000. In this Reynolds number range, normal force

coefficients exhibit an increase just before leading edge

flow separation occurs. This increase in CN may be

attributed to various flow phenomena such as flow reversal,

three-dimensionality of the flow, or to blockage effects

which may be quite significant at such high angles of

attack. The three-dimensionality of the flow alters the

pressure distribution along the span, and is normally

initiated at the corners between the model and the tunnel

walls. These flow phenomena become most evident at high

incidence, as shown in figures 4.3.7 to 4.3.13.

Throughout all the figures of CN and CMc/4 discussed

above, there is an obvious "kink", mentioned earlier. Close

examination of figures 4.4.1 to 4.4.15, may reveal that the

"kink" occurs at angles of attack between -0.10 0 and 3.90 0 ,

depending on Reynolds numbers. Examination of the relative

pressure distributions (Figs 4.1.1 to 4.1.15) did not

indicate any obvious cause of the effect. Previous

investigations, on aerofoil sections at low Reynolds

-59-
numbers, have also shown the existence of "kinks" similar

to those observed here (Refs 26,36,37 and 38). Some of the

results obtained from these investigations are presented in


figures 4.4.16, 17, 18 and 19.

In the present investigation, it was originally


assumed that the "kink" was initiated by the disappearance

of the lower surface separation bubble. However, from a

later analysis of the pressure distribution plots, it was

apparent that this was not the case. The reason for this

is, that, for flow conditions at which the separation

bubble on the lower surface should have disappeared were

not coincident with that of the "kink". Flow visualisation

of the lower surface may have provided information about

the behaviour of the leading edge bubble to support the


above argument.

Further investigation was therefore carried out using

the two-dimensional viscid-inviscid aerofoil analysis

method of Coton and Galbraith (Refs 27 and 39). The outcome

of this work revealed that, for Reynolds numbers greater


than 100,000, the lower surface boundary layer experienced

laminar flow separation well after the incidence at which


the lower surface bubble should have collapsed. Thi s

separation moved slowly towards the trailing edge with


increasing incidence. However, when a certain angle of

attack was reached, the separation point suddenly advanced

to the trailing edge. This implied that the lower surface

-60-
boundary layer, prior to that incidence, was experiencing a

significant region of turbulent flow, whereas, above this

angle of attack, the boundary layer was laminar and fully

attached. A comparison between this angle of attack and the

one at which the "kink" was observed to occur, showed good

agreement (Fig.4.4.20 and Table 4) .

Similar flow phenomena over the lower surface of a

Wortman FX-63-147 aerofoil were also observed by Bastedo

(Ref.33). He claims, however, that the non-linearity of the

CL and CD curve slopes (Fig. 4.4.21) at moderate incidences

was caused by the formation of a separation bubble on the

upper surface. Similar statement was also given by Poll et

al (Ref.37) who states, that the occurrence of the "kink"

on the aerodynamic curves is not only dependent on the

Reynolds number but also to the establishment of the short

separation bubble on the upper surface.

4.5 ~Nmax VARIATION WITH REYNOLDS NUMBERS

This investigation showed a variation in the magnitude

of CNmax at three different ranges, over the considered

Reynolds number range (1. Ox105~Rec~6. Ox10 5 ), see figure

4.5.1. This variation of CNmax is different from that of

the existing data for the same aerofoil tested at higher or

similar Reynolds number ranges (Refs 29 and 34). From these

data, it may be noticed that CNmax rises with increasing

-61-
Reynolds numbers (Figs 4.5.2 and 4.5.3). Further comparison
with data from Reference 28, however, indicates similar-

ities between results (Fig.4.5.4).

Between 100,000 and 200,000 Reynolds numbers, CNmax

occurs at 12.40 0 for the lower Reynolds number regime, but

at 12.90 0 for the higher; as shown in Table 4. The value of

CNmax remains relatively constant at 1.38, with a small

increase to 1.39 at 200,000 Reynolds number.

In the Reynolds number range 2 .5x10 5 <Re c <6. Ox10 5

(Figure 4.5.1), it may be observed that CNmax reduces

steadily from 1.39 to 1.32, with its sharpest reduction

occurring between 400,000 and 450,000. After 450,000,

however, CNmax stabilises at a value of 1.32 whilst the

CNmax incidence reduces to 11.90 0 , where it remained (Table

3). This downward shift is a consequence of an early

initiation of trailing edge separation but subsequent

slower penetration towards the leading edge and the

bursting of the laminar separation bubble (i.e., abrupt

stall) was delayed until several degrees after CNmax had

occurred (Figures 4.5.5(a) and (b)).

From a Reynolds number of 150,000 onwards and with


increased incidence, the movement of the turbulent

separation point appears to reduce its contribution to the

-62-
normal force of the upper surface, while the same increase

in incidence expands the lower surface normal force

contribution. The balancing of these two effects produced a

"flat" region on the CN curve slope prior to complete

stall. Data from References 29 and 26 relating to the

NACA-4415 and Wortmann FX-63-137 aerofoil sections

respectively, shows the same tendency (F igs 4.5.2 and

4.5.6) .

4.6 COMPARISON OF NACA-4415 AEROFOIL SECTION


CHARACTERISTICS WITH EXISTING DATA

Two-dimensional data for the NACA-4415 aerofoil

section have been obtained in the past through experimental

investigations performed in the united States (1937-47) and

West Germany (1962-72), (Refs 28, and 29). The wind tunnels

used for these tests were the NACA LTT (Low Turbulence

Tunnel) and the lAG Stuttgart wind tunnel. Two-dimensional

tests were performed using smooth NACA-4415 aerofoil models

at Reynolds numbers ranging from O.7x10 6 to 3.0x10 6 (Figs

4.5.2 and 4.5.3) These results were corrected of blockage

and streamline curvature effects.

For the comparative analysis between the present and

existing data, extra information was essential about the

type and size of the cross-sectional area of the working

section of the above wind tunnels; the turbulence intensity

together with the model sizes. This information became

-63-
available from References 28, 29, 30 and 31 and is
presented in Table 5.

Since the Reynolds number range at which the NACA LTT

and the Stuttgart wind tunnels were operating was between

0.7x10 6 and 3.0x10 6 , higher than that of the present

investigation (0.5x104~Rec~0.6x105), a direct comparison of

these results was not possible. Therefore, a brief

correlation of the aerodynamic characteristics of these

data was made here, as discussed below.

Variations of normal force and pitching moment

coefficients obtained in this study at a Reynolds number of

600,000, are plotted in figure 4.6(a) against the

corresponding coefficients from the other investigations at

a Reynolds number of 700,000 (Refs 28, and 29). It is clear

from the figure, that all the data are in reasonable

agreement up to an incidence of 110. For angles of attack

greater than this, however, the curves diverge and, in

particular, when maximum lifts are attained at 12 0 , 14 0 and

16 0 they are well separated from each other. The maximum


values of lift coefficients at these incidences were 1.27,

1. 28 and 1. 38 .

From the pitching moment coefficient data (Fig.

4.6(b)) it may be observed that the trends of the curves

are quite similar. Comparing the present data with the

Stuttgart data, the present data exhibit considerably

-64-
higher pitching moments for the entire incidence range.

This difference varies from approximately 0.010 to 0.027

with increasing incidence. In comparison with the NACA LTT

data, however, it may be noticed that the values agree

reasonably well between -2 0 and 60 of angles of attack. For

greater incidences the present data is substantially

higher, especially near the stall.

These variations, observed in the lift and pitching

moment curves, are possibly due to the differences in

turbulence intensity levels as well as the effects of noise

and mechanical vibrations, which in this study are believed

to have being quite considerable. Also, bearing in mind

that the considerable differences between wind tunnels and

model sizes, shown in Table 4, the above variations in the

curves were expected. It was very difficult to predict,

however, exactly how these differences would influence the

results.

4.7 COMPARISON OF NACA-4415 AEROFOIL SECTION


CHARACTERISTICS WITH GU25-5(11)8. NASA-GA(W)-l AND

NACA-0015 SECTIONS

In order to obtain a better overall picture of how

the NACA-4415 aerofoil performs with increasing Reynolds

numbers (Re) and increasing incidence (a), CN contours were

plotted, against the above parameters and are illustrated

-65-
in Figure 4.7.1. The main advantage of the development of

CL or CN contours against a base of incidence and Reynolds

number, is that it shows clearly the overall behaviour of

an aerofoil at various untested Reynolds numbers. Regions

in which the contour lines lie very close to one another,

at a small range of Reynolds numbers, are regarded as

"critical", since the slightest increase or decrease in

Reynolds number or incidence will result in considerable

changes in lift or normal force coefficients.

From Figure 4.7.1, it is evident that the performance

of the aerofoil improves considerably between 50,000 and

75,000 Reynolds numbers. At the lower value, the aerofoil

appears to be in the subcritical range in which it suffers

badly from laminar separation without subsequent

re-attachment. This is demonstrated by the corresponding

pressure distribution behaviour given in Figure 4.1.1, and

by the normal force curve which approximates that of a flat

plate (Fig. 4.4.1) At the higher Reynolds number, however,

it is indicated that the aerofoil operates above the

critical Reynolds number, since its performance is

considerably improved (Fig. 4.4.2).

Figure 4.7.1 also shows that for further increases in

Reynolds numbers up to 450,000, the CN varies very little

for angles of attack less than 10 0 , and is almost constant

for greater incidences and Reynolds numbers. The maximum

normal force coefficient decreases with increasing Reynolds

-66-
number by approximately 5% between 75,000 and 600,000. It

may also be observed from the CN contours that, decreasing

the Reynolds numbers results in reduction of the zero

normal force coefficient angle of attack, from its designed

value of -4 0 to -1.5 0 . This is mainly due to the thickening

of the boundary layer which reduces the effective camber of

the aerofoil at lower speeds.

It could be said, that the overall performance of


this aerofoil is satisfactory, especially for angles of

attack less than 10 0 and for Reynolds numbers greater than

75, 000, indicating that the aerofoil could be safely

operated between these Reynolds number and incidence

ranges.

The comparison of the NACA-4415 with the GU25-5(11)8


(Ref.17) and GA(W)-l, NACA-0015 sections (Ref.18) is

presented in figures 4.7.2, 3 and 4. For this purpose, the

present model was constructed in such a way as to match the

dimensions of the other three aerofoil sections (i. e.

thickness/chord ratio and aspect ratio). The main

assumption of carrying out the above comparison was that if

two or more aerofoil sections having similar dimensions,

being tested under the same wind tunnel environmental


conditions (i.e. turbulence intensity, noise and mechanical
vibrations levels) and using the same wind tunnel

facilities, the results of those tests could be directly

comparable.

-67-
It may be noticed, from the above figures, that the

performance of the GU25-5(11)8 section is strongly affected

by laminar separation without re-attachment, particularly

at Reynolds numbers lower than 350,000. For greater

Reynolds numbers, however, where flow re-attachment occurs,

the performance of this aerofoil increases considerably to

quite impressive standards. This can be seen clearly in

Figure 4.7.2, where, for example, at an incidence of 6 0 and

in a short Reynolds number range between 2. 5x1 0 5 and

4.0x10 5 , the CL magnitude changes dramatically from about

0.35 to 1.17, an increase of more than 240%.

From Figures 4.7.1, 3 and 4, it may also be observed,

that the performance of the GA(W)-l improves at a slower

rate than the NACA-4415 and NACA-0015, between the range of

0.5x105~Rec~1.5x105 and a<5°. For these aerofoils, further

increases in incidence showed little variation in CN until

CNmax was obtained. As incidence and Reynolds number

increased even further, a reduction in CN occurred. This

decrease is more abrupt for the GA(W) -1 and NACA-0015

aerofoils, indicating a leading edge type of stall,

regardless of Reynolds number. For the NACA-4415 section,

however, the loss of lift was more gradual due to trailing

edge flow separation, with the abrupt loss occurring at a

higher incidence.

-68-
As a concluding remark, it may be said that all

aerofoils examined in this investigation appear to have

their own operational limits. The NACA-4415 together with

the GU25-5 (11) 8 aerofoil could be safely operated at

Reynolds numbers greater than 100,000 and 400,000

respectively. The operational limits of the GA(W)-l and

NACA-0015 aerofoils are similar to the NACA-4415 but due to

their early leading edge stall characteristic, they have

been shortened. The contours also indicate that the


GU25-5(11)8 aerofoil stalls at an earlier incidence, while

for the other three aerofoils stall is delayed as Reynolds

number increases. The occurrence of earlier stalling, as

stated by Laing and Kokkodis (Refs 35 and 18), is due to

earlier trailing edge separation caused by the formation of

the separation bubble at mid-chord of the aerofoil. It

should also be mentioned, that maximum normal force

coefficient for the GA(W)-l and NACA-0015 sections

increases slightly with increasing Reynolds numbers, while

for the NACA-4415 and GU25-5(11) 8, it decreases.

The NACA-4415 appears to be, generally, superior to

GA(W)-l and NACA-0015 since it produces higher CN values

for the same increase in incidence and it has more

favourable stalling characteristics. It is also more

favourable to the GU25-5(11)8, but only at Reynolds numbers

lower than 350,000. For higher Reynolds numbers, however,

the GU25-5(11) 8 aerofoil section shows its great

-69-
superiority by producing the highest eN values of all those

sections considered.

-70-
CHAPTER V

CONCLUSIONS AND RECOMMENDATIONS FOR

FURTHER STUDIES

The present study could be described as a preliminary

investigation into the performance of a NACA-4415 aerofoil

section tested at Reynolds numbers ranging from 50,000 to

600,000. The main objectives were to accumulate sufficient

chordwise static pressure data to attain the aerodynamic

forces and moments, namely CN' CT' and CMc/4' and to assess

how these forces and moments have been affected by the

behaviour of the boundary layer around the aerofoil. The

boundary layer behaviour was examined by studying the

static pressure distributions that occurred over the

aerofoil in conjunction with a selection of flow

visualisation photographs which were taken to record the

flow phenomena that occurred over the upper surface of the

aerofoil only.

Furthermore, the aerodynamic characteristics of the

present aerofoil section were compared with those obtained

for the GU25-5(11)8, NASA GA(W)-l and NACA-0015 previously

tested using the same facilities.

-71-
5.1 SUMMARY OF CONCLUSIONS

The analysis of the results from the chordwise

pressure distribution measurements showed that the upper

and lower surfaces of the aerofoil were dominated by the

development of two laminar separation bubbles. The lower

surface bubble, defined as "short", initially forms close

to the leading edge at negative angles of attack. As

incidence is gradually increased it progresses towards the


trailing edge, while its length slowly increases. At a

certain incidence, however, depending on the Reynolds


number, the bubble disappears. In contrast with the lower

surface bubble, that on the upper surface, originally

described as "long", travels rapidly upstream becoming

shorter in length with increasing incidence and Reynolds

number. As it moves to within 15% of the chord from the

leading edge its rate of movement was shown to slow down

considerably. Flow separation was also observed to occur,

sometimes covering an area as high as 65% of the chord from

the trailing edge. When a certain angle of attack was

reached the bubble "burst" causing massive flow separation

over the aerofoil.

The sensi ti vi ty of the boundary layer was also

highlighted during both the pressure and flow visualisation


measurements. It was observed, that at incidences just

prior to complete flow separation the boundary layer

behaved rather erratically. In that it would alternate from

-72-
being fully attached to fully separated, in a non-periodic

fashion within a few seconds. This was thought to be caused

by the large environmental disturbances, at such

incidences, present during the experiments, causing the

boundary layer to flick-on and off from one state to


another.

From a series of oil-film flow visualisation

experiments, conducted to study the nature and behaviour of

the flow over the upper surface of the model, a number of


flow visualisation photographs were taken. In the analysis

that followed, vital information was obtained about the

behaviour of the boundary layer. It was revealed that at

low incidences a "long" laminar separation bubble was

formed close to the trailing edge, moving upstream and


becoming shorter in length as the incidence and Reynolds

number were increased. This gave confirmation of the

existence and behaviour of such a bubble, indicated by the

earlier pressure distribution measurements. Three-dimensi-

onali ty of the flow was also observed, especially for

incidences higher than 11.90 0 and prior to complete stall.

The oil-film flow visualisation technique became very

useful, especially in cases where the laminar separation

bubble, although present, was unable to be detected by the


pressure measurement method. However, when detection of the

separation bubble and trailing edge flow separation was

obtained, there was generally a close agreement between the

-73-
results obtained by the two methods; in particular, at a

Reynolds number equal to 600,000. For the lower Reynolds

number ranges, there was some slight disagreement between

the laminar and turbulent flow separation points, while

turbulent re-attachment points matched each other exceptio-

nally well for all incidences and Reynolds numbers.

The aerodynamic forces and moments generated to study

the performance of the aerofoil in these low Reynolds

number regimes, were found also to be affected by the

behaviour of the boundary layer. In cases where the

development and disappearance of the separation bubbles had

occurred, on either the upper or lower surface of the

aerofoil, dCN / da and dCMc / 4 / da were affected. This

manifested itself in an obvious magnitudes change. This was

most evident when the lower surface turbulent boundary

layer changed to fully laminar state over a narrow range of

incidences. As a result of this, a slight "kink" developed

on the aerodynamic coefficient curves and was present over

the majority of the Reynolds numbers tested.

The most dominant feature in the normal force and

pitching moment curves, apart from the "kink", was the

gentle decrease in CN and an increase in CMc/4 and CT with

increased incidence. This was due to the obvious

penetration of the trailing edge separation towards the

leading edge. When complete stall occurred an abrupt

-74-
reduction in aerodynamic coe fficient magnitudes was

observed. From an examination of the pressure profiles,

this was caused by the "bursting" of the upper surface

separation bubble in the region close to the leading edge.

Therefore, the behaviour of the normal force and the

quarter chord pitching moment coefficients, suggested that

the stalling characteristic of the NACA-4415 aerofoil

section was that of a "combined stall", i.e. a combination

of trailing and leading edge stall. As Reynolds number was

increased, stalling of the aerofoil was found to occur at a

higher incidence. The maximum normal force coefficient also

varied for Reynolds numbers below 400, 000. For higher

Reynolds numbers, however, it remained relatively constant

having a value of approximately 1.32.

The critical Reynolds number for the NACA-4415

aerofoil was found to lie between 50,000 and 75,000. It was

demonstrated that, when the aerofoil operated below the

critical Reynolds number, its performance was similar to

that of a flat plate, while when operated above it, it

behaved like a conventional aerofoil. Although the aerofoil

was well behaved for Reynolds numbers between 450,000 and

600,000, its performance below this regime was more

dependent on Reynolds numbers, especially at incidences

greater than 8 0 .

Present and existing data were also examined and

showed that the differences in test environments, wind

-75-
tunnels, model sizes, etc., can have considerable effects.

With this in mind, however, it is difficult to predict the

way they will affect the results of this particular series


of tests.

Lastly, comparing the characteristics of the present

aerofoil with those of the GU25-5(11)8, NASA GA(W)-l and

NACA-0015, revealed that the GU25-5(11)8 showed superiority

over the other three aerofoils for Reynolds numbers above

350,000, by producing the highest CN values. For lower

Reynolds numbers, however, it behaved in a very unstable

manner, with all the other aerofoils greatly superior to

it, especially at moderate to low incidences. The NACA-4415

aerofoil is pre-eminent over GA(W)-l and NACA-0015 for all

Reynolds numbers tested, as it generates higher CN values,

has better stalling characteristics and a higher

operational range.

5.2 RECOMMENDATIONS FOR FURTHER STUDY

The NACA-4415 aerofoil section is known to be widely

used on the rotor blades of various Horizontal Axis Wind

Turbines (HAWT). These turbines operate in an open

atmospheric environment and therefore are exposed to the

elements. Because of their operational environment some

roughness to the blades in the form of rain droplets, dust


or even in the form of bird droppings and dead insects, may

exists. These can affect the performance of the blades and

consequently that of the turbine. Bearing in mind such

-76-
condit ions , it is therefore suggested that further

experiments should be carried out with the application of

some roughness on either the upper or lower surface of the

present NACA-4415 aerofoil model. The applied roughness

could be in the form of a trip wire covering the whole

length of the span, having different diameter sizes,

located at various positions along the chord length, etc.

As mentioned in Section 4.4, the existence of the

slight "kink" on the CN and CMc/4 curve slopes was thought

to have been caused by the occurrence of full laminar flow

on the lower surface of the aerofoil. It is therefore

advisable, that in future, a surface oil-film flow

visualisation technique should be employed to study the

boundary layer behaviour on the lower surface. Its use

might prove most useful to confirm or refute the sudden

downstream movement of the laminar separation point

inferred by using the method of Coton and Galbraith (Refs

27 and 39). Additionally, since the surface oil flow

visualisation technique, utilised for studying the

behaviour of the boundary layer over the upper surface of

the model, gave apparently misleading results for Reynolds

numbers below 200,000, an alternative technique, such as

the direct smoke injection, could be recommended.

For a limited number of experiments, however, it was

noticed that the upper surface separation bubble, although

present, as indicated by the flow visualisation pictures,

-77-
could not be identified from the pressure distribution

plots. In such cases, further experiments are required to

examine the nature and behaviour of the flow. This could be

accomplished by using a large number of hot film anemometry

gauges over the upper surface of the model.

Finally, due to the large amount of time consumed in

the present investigation for the data acquisition

procedure, an alternative method should be employed and a

possible alternative is described below.

The pressure tappings used for recording the pressure

variation around the present aerofoil at various incidences

and Reynolds numbers, should be substituted by a number of

miniature pressure transducers. By employing such pressure

transducers an automatic and simultaneous pressure

recording can be obtained with the help of a Transient

Recording System linked to a highly compatible computer.

The computer could be programmed to display the pressure

distribution around the aerofoil in a graphical form on its

visual display terminal or onto a printer after each test

run. Such a procedure will allow the user to check for any

faults in the collected data immediately after each test

and if necessary to repeat the test. When all test runs are

completed, the computer can perform the necessary integra-

tions of each pressure distribution, producing plots of the

aerodynamic forces and moments. This will enable the user

to start an immediate analysis of the aerodynamic chara-

-78-
cteristics of that particular aerofoil.

Such a data acquisition and data reduction procedure

could not only save a considerable amount of the

researcher's time, but would allow him also to test more

aerofoils inside the permitted time given. The shortcomings

and difficulties for the employment of such a method,


however, are recognised.

5.3 CONCLUDING REMARKS

The results of this study provided valuable

information about the aerodynamic characteristics of the

NACA-4415 aerofoil and the behaviour of the boundary layer,

when tested two-dimensionally for Reynolds numbers below

600,000. From the acquired data, speculations may arise

about the validity of the results, mainly because of the

high levels of free stream turbulence, noise and mechanical

vibrations in the working section of the test facility.

However, because the above aerofoil is used extensively in

HAWT applications, which in turn are operating in an

environment possessing similar perturbations, the present

data could be considered as valid.

A search of the available literature did not bring to


light any two-dimensional data about the NACA-4415 aerofoil

tested at Reynolds numbers below 700,000. Therefore, since

the blades of HAWT machines, designed and constructed for

-79-
wind tunnel experimental purposes, operate usually in

regimes lower than 1,000,000, the present data could prove

very valuable.

Finally, due to the uncertainty of the performance of

aerofoil sections tested below Reynolds numbers of 500,000,

as shown in the present investigation, it is hoped that

this will encourage future researchers to examine the

behaviour of other aerofoils currently operating in these

low Reynolds number regimes.

-80-
APPENDIX

MANUFACTURE AND ASSEMBLY OF NACA-4415 AEROFOIL MODEL

The construction of the NACA-4415 aero foil model was

carried out in the manufacturing facilities located in the

Aerospace Engineering Department of the University of

Glasgow. The facilities consist of two blocks of wax

together with a fixed head cutting machine, fitted with

router and follower which was used to cut the wax into the

required aerofoil shape. Figure A1 shows the model under

construction, divided into two halves, as well as the

locations of the structural materials used.

Before the cutting of the wax, the x and y

coordinates of the NACA-4415 aerofoil were plotted onto a

1/4 inch thick mild steel plate using a vernier height

gauge to ensure accuracy of the profile. The steel plate

was cut to provide the required upper and lower surfaces of

the aerofoil and mounted on the wax cutting machine with

the router follower resting on it. After machining, Slipwax

was applied to ease mould release and the wax surface

polished to give a good finish. Epoxy resin gel-coat was

then spread and left to harden. This gel-coat covered both

surfaces for the required span length of the model.

-81-
Holes representing the pressure orifices were drilled

through the hardened surface of the gel-coat into the wax

at the chosen surface locations. At this stage, pins having

the same diameter were inserted into the pressure orifices

preventing them from getting blocked when glass fibre and

epoxy resin were added to model. Four layers of 280 gm/m2

glass fibre woven roving together with epoxy resin applied

between each layer were laid on the gel-coat surface at an

angle of +/- 45° to give a strong torsional stiffness.

Balsa wood blocks were then cut to fill the aerofoil

shape and positioned at each end of the model, allowing

room for 1/4 inch aluminium plates to finish off the ends.

The relevant balsa wood was then drilled to allow for the

extension of the brass pressure tubes out past the length

of the model. When these blocks were in place the brass

pressure tubes were allocated on the model. Epoxy resin

foam mixture was then poured into the model and left to

harden. When this was complete the model was ready to be

machined flat. After machining, either the upper or lower

surface model was removed from the wax block and placed on

top of the other so that the leading edges and balsa wood

blocks were parallel with each other. While in position,

four holes were drilled through the balsa woods to take 3/8

inch dia. dowel pins. The pins prevented either half of the

model from slipping. Epoxy resin mixture was then applied

to both flat surfaces. The top half of the model was then

pressed carefully to squeeze out any excess foam. Extra

-82-
care was taken by checking constantly that no slipping

movement had occurred since the slightest movement could

misalign the whole model.

When the resin had set, the whole model was removed

from the wax mould. The aluminium end plates were filled to

the ends of the model and fixed by woodscrewing them on the

balsa wood blocks in conjunction with an epoxy bond. The

accuracy of the model's profile was carefully checked using

the templates and bumps or indentations on the surface were

rectified by filling in with gel-coat and rubbing down with

wet and dry emery paper. The brass pressure tubes were

checked for any blockages and cleaned out by blowing

through with compressed air. Finally, when accuracy of the

aerofoil profile was achieved and the pressure tubes

cleaned, the model was ready to be installed into the

tunnel. The connection of the brass pressure tubes with the

pressure ports at the rear of the selector boxes was

achieved using plastic tubes.

-83-
REFERENCES

1. Mueller, T.J.
"Low Reynolds number Vehicles" AGARDOgraph No.288, February
1985.

2. Mueller, T.J., Jansen, B.J.Jr.


"Aerodynamic Measurements at Low Reynolds Numbers", AIAA
paper No. 82-0598, 1982.

3. Marchman, J.F.
"Aerodynamic Testing at Low Reynolds Numbers", Journal of
Aircraft, Vol.24, No.2, February 1987.

4. Lissaman, P.B.S.
"Low Reyn olds Number Air foils", Annual Review of Fluid
Mechanics, Palo Alto, California: Annual Reviews, Inc., Vol.
15 , pp . 223 - 23 9 , 1 98 3 .

5. Pope, A., Rae, W.H., Jr.


"Low Speed Wind Tunnel Testing", John Wiley & Sons, New York,
1984.

6. E.S.D.U.
"wind Tunnel Corrections", 'Aerodynamics', Volume 13, Item
Number 76028, November 1976.

7. Mueller,T.J., Pohlen,L.J., Conigliaro,P.E, Jansen,B.J.Jr.


"The Influence of Free Stream Disturbances on Low Reynolds
Number Airfoil Experiments", Experiments in Fluids, Vol.1,
pp. 3-14, 1983.

8. Sumantran, V., Sun, Z., Marchman, J.F.III.


"Acoustic and Turbulence Influence on Low Reynolds Number
Wing Pressure Distributions", Conference on Low Reynolds
Number Airfoil Aerodynamics, Notre Dame, June 16-18, 1985.

-84 -
9. Mueller, T.J.
"The Influence of Laminar Separation and Transition on Low
Reynolds Number Airfoil Hysteresis", AIAA 17th Conference on
Fluid Dynamics, Plasma Dynamics and Lasers, Colorado, June
25-27, 1984.

10. Horton, H.P.


"Laminar Separation Bubbles in Two and Three Dimensional
Incompressible Flow", Ph.D. Thesis, Queen Mary College,
University of London, London, 1968.

11. Chappell, P.D.


"Flow Separation and Stall Characteristics of Plane, Constant
-Sections Wings in Subcritical Flow", The Aeronautical
Journal of the Royal Aeronautical Society, Vol. 72, pp.
82-90, January 1968.

12. Arena, A.V., Mueller, T.J.


"Laminar Separation, Transition, and Turbulent Re-attachment
near the Leading Edge of Airfoils", AIAA Journal, Vol. 18,
No.17, pp.747-753, July 1980.

13. Gaster, M.
"The Structure and Behaviour of Laminar Separation Bubbles",
ARC R & M, No. 3595, March 1969.

14. van den Berg, B.


"Role of Laminar Separation Bubbles in Airfoil Leading-Edge
Stalls", AIAA Journal, Vol.19, No.5, pp. 553-556, May 1981.

15. Ward, J.W.


"The Behaviour and Effects of Laminar Separation Bubbles on
Aerofoils in Incompressible Flow", Journal of the Royal
Aeronautical Society, Vol.67, pp. 783-789, December 1963.

-85 -
16. Kokkalis, A.
"An Investigation into the Wake Flow Behind Circular
Cylinders Fi tted wi th Slat Devices", M. Sc. Thesis, Department
of Aeronautics and Fluid Mechanics, University of Glasgow,
March 1983.

17. Lunde, K.
"Wind Tunnel Testing of Aerofoil Sections", Final Year
Project, Department of Aeronautics and Fluid Mechanics,
University of Glasgow, April 1983.

18. Kokkodis, G.
"Low Reynolds Number Performance of a NACA-0015 and a GA (W)-1
Aerofoil", M.Sc. Thesis, Department of Aeronautics and Fluid
Mechanics, University of Glasgow, February 1987.

19. Galbraith, R.A.McD, Coton,F.N., Saliveros,E., Kokkodis,G.


"Aerofoil Scale Effects and the Relevance to Wind Turbines",
Ninth Annual Wind Energy Conference, Edinburgh, Scotland, 1-3
April, 1987.

20. Spalding, D.J.


"Wind Tunnel Testing of Aerofoil Sections", Final Year
Pro j ect, Department 0 f Aeronaut ics and Fluid Mechanics,
University of Glasgow, March 1985.

21. Kelling, F.B.


"Experimental Investigation of a High Lift Low Drag
Aerofoil", G.U. Report No. 6802, Department of Aeronautics
and Fluid Mechanics, University of Glasgow, 1968.

22. Furness Controls Limited.


"Instruction Manuals of FC002 and FC012 Micromanometers".

23. Duncan, W.J., Thorn, A.S., Young, A.D.


"Mechanics of Fluids", Published by Edward Arnold Ltd.,
Second Edition, London, 1970.

-86 -
24. Houghton, E.L., Carruthers, N.B.
"Aerodynamics for Engineering Students", Published by Edward
Arnold Ltd., Third Edition, London, 1982.

25. Saliveros, E.
"Collection of Flow Visualisation Photographs over the Upper
Surface of a NACA-4415 Aerofoil", Report in Progress,
Department of Aerospace Engineering, University of Glasgow,
February 1988.

26. Huber, A.F.II.


"The Effects on Roughness on an Airfoil at Low Reynolds
Numbers", M.Sc. Thesis, University of Notre Dame, Notre Dame,
Indiana, May 1985.

27. Coton, F.N.


"Contribution to the Prediction of Low Reynolds Number
Aerofoil Performance" f Ph.D. Thesis, Department of
Aeronautics and Fluid Mechanics, University of Glasgow,
December 1987.

28. Loftin, L.K., Jr., Smith, H.A.


"Aerodynamic Characteristics of 15 NACA Airfoil Sections at
Seven Reynolds Numbers From O.7x10 6 to 9. Ox10 6 ", NACA
Technical Note 1945, October, 1949.

29. Althaus, D.
"Stuttgarter Profilikatalog I", Institut fur Aerodynamik und
Gasdynamik der Universitat Stuttgart, 1972.

30. Miley, S.J.


"A Catalog of Low Reynolds Number Airfoil Data for Wind
Turbine Applications", Department of Aerospace Engineering,
Texas A&M University, Texas 77843, February 1982.

31. Riegels, F.N.


"Aerofoil Se ct ions", Publi shed by But terworths and Co.
(Publishers) Ltd., London, 1961.

-87 -
32. Galbraith, R.A.McD.
"The Aerodynamic Characteristics of a GU25-5 (11) 8 Aerofoil
for Low Reynolds Numbers", Experiments in Fluids, Vol.3, pp.
253-256, 1985.

33. Bastedo, W.G.Jr.


"Performance of an Aerofoil and Three Rectangular Planform
Wings at Low Reynolds Numbers", M.Sc. Thesis, Department of
Aerospace and Mechanical Engineering, Notre Dame, Indiana,
December 1984.

34. Jacobs, E.N., Sherman, A.


"Airfoil Section Characteristics as Affected by Variations of
the Reynolds Numbers", NACA Technical Report No. 586, 23rd
Annual Report, 1937.

35. Laing, S.
"Wind Tunnel Test ing of Aerofoil Sect ions", Final Year
Pro ject, Department of Aeronautics and Fluid Mechanics,
University of Glasgow, April 1984.

36. Clayton, B.R.


"BWEA Initiative of Wind Assisted Ship Propulsion", Journal
of Wind Engineering and Industrial Aerodynamics, Vol. 19, pp.
251-276, 1985.

37. Poll, D.I.A., Mansoor, A.


"On the Determination of the Two-Dimensional Characteristics
of Aerofoils at Low Reynolds Numbers", International
Conference on Aerodynamics at Low Reynolds Numbers
(10 4 <Re<10 6 ), London, 15-18 October, 1986.

38. van Ingen, J.L., Boermans, L.M.M.


"Aerodynamics at Low Reynolds Numbers: A Review of
Theoretical and Experimental Research at Delft University of
Technology", International Conference on Aerodynamics at Low
Reynolds Numbers (10 4 <Re<10 6 ), London, 15-18 October, 1986.

-88 -
39. Coton, F.N., Galbraith, R.A.McD.
"A Simple Method for the Prediction of Separation Bubble
formation on Aerofoils at Low Reynolds Numbers",
International Conference on Aerodynamics at Low Reynolds
Numbers (104<Re<10 6 ), London, 15-18 October, 1986.

-89 -
TABLE 1.
Coordinates of NACA-4415 Aerofoil Section

UPPER SURFACE LOWER SURFACE


x/c y/c x/c y/c

1. 00000 0.00156 0.99979 -0.00156


0.96946 0.01100 0.96771 -0.00284
0.93865 0.02002 0.93576 -0.00415
0.90781 0.02863 0.90397 -0.00505
0.87697 0.03682 0.87236 -0.00689
0.84617 0.04460 0.84095 -0.00833
0.81545 0.05198 0.80978 -0.00981
0.78483 0.05814 0.77887 -0.01134
0.75437 0.06550 0.74824 -0.01292
0.72408 0.07164 0.71793 -0.01454
0.69400 0.07738 0.68795 -0.01619
0.66418 0.08270 0.65834 -0.01787
0.63463 0.08762 0.62911 -0.01956
0.60540 0.09211 0.60029 -0.02126
0.57652 0.09620 0.57191 -0.02294
0.54802 0.09986 0.54399 -0.02460
0.51994 0.10311 0.51654 -0.02622
0.49231 0.10594 0.48960 -0.02778
0.46515 0.10835 0.46319 -0.02926
0.43851 0.11034 0.43732 -0.03066
0.41241 0.11192 0.41201 -0.03195
0.38662 0.11306 0.38756 -0.03314
0.36119 0.11363 0.36396 -0.03433
0.33640 0.11364 0.34097 -0.03552
0.31228 0.11310 0.31861 -0.03667
0.28888 0.11203 0.29690 -0.03776
0.26621 0.11046 0.27584 -0.03878
0.24433 0.10842 0.25544 -0.03968
0.22324 0.10592 0.23572 -0.04046
0.20299 0.10302 0.21669 -0.04109
0.18360 0.09972 0.19836 -0.04156
0.16508 0.09607 0.18074 -0.04185
0.14748 0.09209 0.16385 -0.04194
0.13080 0.08782 0.14770 -0.04181
0.11507 0.08329 0.13231 -0.04147
0.10030 0.07854 0.11768 -0.04088
0.08651 0.07360 0.10383 -0.04006
0.07370 0.06850 0.09078 -0.03898
0.06190 0.06327 0.07854 -0.03765
0.05110 0.05794 0.06712 -0.03604
0.04133 0.05255 0.05655 -0.03417
0.03257 0.04712 0.04683 -0.03203
0.02484 0.04168 0.03798 -0.02960
0.01814 0.03624 0.03001 -0.02690
0.01247 0.03084 0.02294 -0.02392
0.00738 0.02549 0.01154 -0.01709
0.00164 0.01501 0.00386 -0.00912
0.00000 0.00000 0.00000 0.00000

-90-
TABLE 2.
Locations of Pressure Tappings on NACA-4415
Aerofoil Section

UPPER SURFACE
x/c y/c x/c y/c

0.979666 0.007933 0.234400 0.108833


0.959866 0.013533 0.193267 0.103267
0.940166 0.018966 0.153267 0.095133
0.919166 0.024833 0.133600 0.090067
0.899200 0.030200 0.114600 0.084633
0.878966 0.035600 0.104533 0.081200
0.860866 0.040066 0.095866 0.078200
0.840266 0.045466 0.085334 0.074466
0.820334 0.050400 0.076000 0.070733
0.799466 0.055200 0.065800 0.066533
0.780000 0.059466 0.055266 0.061267
0.739466 0.068072 0.045733 0.056266
0.679266 0.079626 0.035200 0.051300
0.619266 0.089691 0.027133 0.042666
0.558733 0.098301 0.015800 0.035800
0.500266 0.105039 0.008733 0.028733
0.437200 0.110399 0.004334 0.025000
0.375533 0.113334 0.000667 0.012466
0.315667 0.113066 0.000025 0.003667
0.275466 0.110977

LOWER SURFACE
x/c y/c x/c y/c
0.001533 -0.009866 0.181033 -0.042166
0.010166 -0.015466 0.234733 -0.040531
0.016766 -0.020533 0.281466 -0.038440
0.028466 -0.026334 0.382000 -0.033394
0.039667 -0.030485 0.480733 -0.028366
0.049600 -0.033766 0.581266 -0.022600
0.059466 -0.036165 0.681266 -0.016833
0.069933 -0.038031 0.779334 -0.011600
0.079466 -0.039370 0.879866 -0.006777
0.100600 -0.041061 0.979800 -0.002425
0.139433 -0.042366

-91-
TABLE 3.
2-D ANGLE OF ATTACK SUMMARY

Reynolds CN =0 C [CMc/4Jmax Stall


Number Nmax

50,000 -1.50 4.90 7.90 -----

75,000 -2.90 12.40 14.90 16.90

100,000 -3.80 12.40 14.40 16.65

125,000 -4.10 12.90 14.90 17.40

150,000 -4.20 12.90 14.90 17.40

175,000 -4.20 12.90 14.90 17.40

200,000 -4.10 12.90 14.90 17.90

250,000 -4.10 11. 90 13.90 18.40

300,000 -4.10 11. 90 13.90 18.90

350,000 -4.10 11.90 13.90 18.90

400,000 -4.10 11.90 13.40 19.40

450,000 -4.10 11.90 13.90 19.40

500,000 -4.10 11. 90 13.90 20.40

550,000 -4.10 11.90 14.90 -----

600,000 -4.10 11.90 14.90 -----

-92-
TABLE 4.
Estimated locations of laminar separation points on the lower
surface of a NACA-4415 aerofoil section using a viscid-inviscid
analysis method.

Rex1o' 100.0 150.0 200.0 250.0 300.0 350.0 400.0 450.0 500.0 550.0 600.0

ex (0) x/c (% )

-2 7.7 7.7 8.0 16.6 16.6 17.6 17.6 17.7 17.7 17.7 17.7

-1 8.9 8.9 21. 0 23.1 23.1 23.1 23.2 23.3 23.4 23.5 23.5
I
~
W
I 0 11. 6 12.6 40.6 40.9 42.0 42.4 42.8 57.0 58.1 54.9 51.1

1 14.3 22.1 87.4 92.0 90.5 85.8 81.0 76.0 72.3 69.1 66.4

2 19.2 95.5 100.0 100.0 100.0 100.0 100.0 97.4 94.8 92.0 90.8

3 38.7 100.0 100.0 100.0 100.0 100.0

4 100.0
---_._.-
TABLE 5.
Useful Information About Past and Present Wind Tunnels Used
to Test NACA-4415 Aerofoil Sections.

Test Velocity Turbulence Chord Aspect


Tunnel Type Test
Section (m/s) Intensity Ratio
(m)

NACA LTT
Closec O. 90mx2 .291
Low Turbulence 70 0.03% 0.6 2-D 3.81
Return (Closed)
Tunnel
I
\.0
.J::,.
I
lAG STUTTGART=l
Closed o. 73mx2 . 73 h 0.5 1 5.46
Stuttgart & &
Laminar
91 0.02% 2-D
Return
(Closed) 1.0
Wind Tunnel 2 2.73

GLASGOW
UNIVERSITY Closed 0.84mx1.14
30 ~0.50% 0.3 2-D 2.80
Low Speed Return
(Closed)
Wind Tunnel
1: for Reynolds numbers up to 1.5x10 6
2: for Reynolds numbers> 1.5x10 6

(Adapted from References 28, 29, 30 and 31)


HIGH AlTlTUD! AIRCR"FT
PLATFORMS AND ~'.lNl - "'PVa
1.0
(C.on tt.I~I>c<lJ

li>
.....
'00
a:
:E
~ 0.1 UJ
C)
~
::> '0 ::>
oUJ z
- 0.01 :x:
W ()
0.. «
1.0
...
W

::t
:E
....
o 0.001 ~
:::; -J
u. 0.1 UL1RALIGHT I,lRCnAFT
U.
MODEL AIRPlAIIES
0.0001

0.01

CHORD REYNOLDS NUMBER, Uc


l!

Figure 1.1. Chord Reynolds Number versus Flight Velocity


for a Variety of Natural and Man-Made Flying
Objects. (Adapted from Reference 1)

Sep. Bubble with turbulent reattachment


Probably local
separation near
trailing edge

Figure 1.2. A Typical Laminar Separation Bubble Formed


over the Upper Surface of an Aerofoil.
(Adapted from Reference 1)

- 95-
CP
A: 0 ~ - 12.5 Ret- IOOeGB .132% UPPER
-5 B: ¢ I( - 12.5 RCf- I emlElB .02% LOHER
c: A " D 12.5 RCt- lomme .2% LO~ER
-4
-3

,,
CJ - ,TUR8uL[1C([ 'nH~ptSITT 0.021

6 - TUHzlUUhC£ IptTI"'~IlT 0.21

Re ... 100.000

-11 II , ze
+tH+~+H+H+H'+'~II~'~II~I'+II~I+II~I+1~II+'HII+'~III

(flaj a (b)

Figure 1.3. Turbulence Effects on Stall Hysteresis (a) and


Pressure Distribution (b) for the Wortmann
FX-63-137 Aerofoil Section. (Re=100,OOO)
(Adapted from Reference 7)
A'-. 6.->. CP
u Ai'~' ~ A: a 16.5 RCf· 2000013 .02%

~A~';"C
IJ a

-5 B: 0 Q 0= 16.5 RCi= 20000e .02%


c: 6. tJ c 16.5 RCf= 2000eIJ .2%
tJ." -4
B-?,
, .'

I
I
I
GJ - Tt;I;,::.Lr... l'i;:( I1ITC.OIjtTY 0.011
A - TI.'1i.2~-... ('1([ !llT[l1SITY 0.2:

,I rtf! ... 200.1Y.lO

I
-II II I 11
+j-I'i31:19+-~r++H+H++-t+H-H-+'H-+ I' I I I r I I I )-

a (b)

Figure 1.4. Turbulence Effects on Stall Hysteresis (a) and


Pressure Distribution (b) for the Wortmann
FX-63-137 'Aerofoil Section. (Re=200,OOO)
(Adapted from Reference 7)

CP
o tJ '" I~.S RCt e 2833013 A
1.:1- A -5 - ¢ a c 14.5 RC~e 20e200 B

\JJ
C A ~ c 14.5 Rn= 2130130e C
-4 -
I
/ I
B'
,:. w ~@:l~

I
CJ - 00 ~CO"SIiCS
Re." lOG.OOO
6, - 5b'lO t-fl.

18 ,18 39
I I II I I I I II I , I' I I I I I I I I I'
(a) a (b)

Figure 1.5. Acoustic Effects on Stall Hysteresis (a) and


Pressure Distribution (b) for the Wortmann
FX-63-137 Aerofoil Section. (Re=200,OOO)
(Adapted from Reference 7)

- 96 -
Dividing Separated Turbulent
Slreamlin~ Shear Layer

Edge of the
Boundary Layer

1..0
-J

Separated
Laminar Shear
Layer "D eo d A'Ir .. Redeveloping Turbulent
Laminar Boundary Layer
Region
Boundary Layer

Figure 1.6. Sketch of a Laminar Separation Bubble.

(Adapted from Reference 10)


/'
/ ,':
/ .' \ ,--Tral'1'lng Ed ge Sta 11
f "
/: , ...... , ~: ______ Combined Stall
/ .
I: ...-/ . :...--' -----:-: ~'--Leading Edge Stall
I .
I..: /
'-"-~ Thin Aerofoil Stall
I.' .
l!
/,~

1/

a;

Figure 1.7. Four Different Types of Static


Stall

Laminar Separation
Transition
- Reattachment
Turbulent Separation

High Angles of Attack


Low Angles of Attack
TS
.. . .
..... 1·0
o .. . ., . ..
x/c

1
Figure 1.8. Typical Aerofoil Pressure Distribution

with Laminar Separation Bubbles.

·98 -
Xl0-l

"
NACA-1115 AeroFoil
45'
Angle of Attack -5. 1 0 (deg.)
I
I
Reynolds Number 298051 .1
40
Dynamic Pressure 131.151 N/m sq.
Uncorrecled Datu

o Upper SurFace
() Lower SurFacE

I
25
- C
P

\-0 20
\-0

10

5'

01 (jfrL; :2 :3 .0\ 5 6~ ~Io


XlO- 1

X/c
-5

Figure 1.9. Experimental Static Pressure Distribution over the NACA-4415


-10
o
Aerofoil at an Incidence of -5.10 and at a Reynolds Number
of 298,051.
straighteners Fan Housing
~
~ '-- I \ i---"" / It--
------
~
~
---
Dl. f fuser
\

Contraction
Diffuser
'\,
I
1
I
t-'
o
o --+--~- II
1
I
I
~--- 1/ 1

Breather Slot ~--------~- Working Section /


Settling Chamber Corner
Vanes

O.383mx1.143m WIND TUNNEL

Figure 2.1. A Plan View of the Glasgow University's Low Speed Wind
Tunnel.
PRESSURE TUBES
GRADUATED ANGLE
SCALE
/
-... /

_ MODEL
1

1.143m

':. PRESSURE

-------- TAPPINGS i
O.838m
_____________ ~_~i______~
i

1\
\
~~ __. -__r -_____________\~\ ____________- r__~__- J

\\
CIRCULAR STEEL
PLATES

/ / I / /7 / / / / / / / / / / / /////7// ////

Figure 2.2. Cross-Sectional View of the Working


Section of the Glasgow University's Low
Speed Wind Tunnel.

-101-
NACA-4415 AEROFOIL SECTION

X10- 2
12

~
o
l'0
10
8
6 ........-------- --------
-2
-4

• •
X/C (X10- 1)
-6

Figure 2.3. Positions of the Pressure Tappings Around the


NACA-4415 Aerofoil Model.
-5 WORTMANN FX 63-137
Rc = 200.000 (c = 6")
Alpha = 16 degrees
Uncorrected data
-4-

"Q..
U
'V -3- [> In-line tap model
0: StaggQred tap model
+)
[
OJ
·rl
U -2-
'rl
f-'
0 4-
w tj-.
(U
0
u -1 .

cu
l ~ e 0
J I>
en
en a
OJ 25 50 75 100
l ~ e ~ e ~ A ~
(L . ..-..- Q A &-()

x/c (7.)
1

Figure 2.4. Pressure Distribution Comparison Between two Wortmann Aerofoils


0
Having Different Pressure Tap Configurations at 16 Angle of
Attack and Re=200,OOO. (Adapted from Reference 1)
-5 T WORTMANN FX 63-137
Rc = 80, 000 (c = 6" )
Alpha = B degroes

I
Uncorrected data
-4-

r-..
(L
U
'---' -3 t I> In-line top rHode 1
0 Staggered top model
~J
r:
ill
rl
u -2
rl
f-' (-t-
o
II'> l~_

ill
0
U -1 -
QJ
L
J
(J)
(J)
m o I -ItA ~ ;5 ;~ S ; ~
+--i
75
t
100
L
(L
Rf¥-
x/c eX)
1

Figure 2.5. Pressure Distribution Comparison Between two Wortmann Aerofoils


Having Different Pressure Tap Configurations at 8 0 Angle of
Attack and Re=80,OOO. (Adapted from Reference 1)
(
START

START TUNNEL &


BRING IT UP TO
DESIRED SPEED
i

RUN PROGRAM
~

AEROFOIL.BAS

CHANGE AUTO LOGGING


INCIDENCE SEQUENCE

END OF PROGRAM

DO
YES YOU
WANT
ANOTHER
RUN

NO

STOP TUNNEL
SPEED

Figure 3.1. Sequence of events during pressure distribution


measurements

-105-
AEROFOIL C ~
I

j l PRESSURE
TAPPINGS

DRIVE
DY1:
DRIVE
DYO:
-
AID
CONVERTER
(MNCAD)
.. -

..-
MICROMANOMETER
I
I FC012 I
J-

I
00
p
jX\

~ SELECTOR - II

~
~ BOXES

........ ... REAL TIME


f-'
o
p
DEC CLOCK III
CJ\ MINC 11/23 (MNCKW)
I I MICROCOMPUTER
.....
p

DEC VT·l05
TERMINAL ,,
~.
.....
... ......
}~
IEEE BUS CONTROLER 488-1975

I I
.1 1
II
PRINTER
IBM EPSON MX-80

Figure 3.2. Schematic of Instrumentation used for the data acquisition.


1.6 6
MILEY

1
lISSAMAN
A,· 150.000
~~
Rc: 150,000 (j:;
1.2 1.
1.2 °0o~
<>

o8L~
<>
c\
0.8
C\
I I 0:>
O(f ~
08 o0:;;«0'),-/
_rA

0
0
o
0° -~
0 00
0.41. 0.4 i 0
v
o <>
~····-t ____ .2....t(L-+ __'-+-_'--""---i ~IO
(>
o
o o
(>
<,:..:.i>-0,4 rJfXXx;O<>
({} -04
0
l--'
0.3
<1
<>
Cd o
<>

J4
o 0.3
-J
Cd

0 ..2 <> (p <>


o 02
00 I8
~;t
00 0
"0, 00 00
~ 0.1 ~o0
0
01 00
cP. 00
0
/'IJl)dY , 0
00 . ,(y 0;;0">
I , '+---'--l~"-"--+-'---i--~ I I .-..-----i~--....- - I 1
-20 -12 -4 4 12 20 :t8 -;1:0 -12 -4 4 12 20 28
a.IOEGR[(SI a.(OECREES)

Figure 3.3. Section Lift and Profile Drag Figure 3.4. Section Lift and Profile Drag
Coefficients Versus Angle of Coefficients Versus Angle of
Attack for Rc=lS0,OOO for the Attack for Rc=lS0,OOO for the
Smooth Lissaman Airfoil. Smooth Miley Airfoil.

(Adapted from Reference 9) (Adapted from Reference 9)


-
o
00
I

Figure 3.5. The NACA-4415 Aerofoil Installed in the Glasgow University's Low
Speed Wind Tunnel
-c
p

-c
p

-1

Figure 4.1.1. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=50,OOO.

-109-
-c

-c
p

-1

Figure 4.1.2. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=75,OOO.

-110-
-c
p

-c
p

Figure 4.1.3. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=100,OOO.

-111-
-c
p

~~~~22.90 21.90
20.90
19.90
18.90
17.90
-c 17.';0
p 16.90
16.';0
15.90
15.+0
' .... 90
13.90
12.90
11.90 e"
"I

9'.~090 ae
7.~090 L1§'
90
5.;0 o-2t
.... 90 'I..J"'v
3.90
2.90 0' "

O.~090 "e (b)


_1~0'0 ,,<'0,
-2.10
-3.10
-".10
'r.,fflt../-~r-T--t---,t--7:---,j---'S---<!9.---=::,-r: -5.10
Xl0- 1

Figure 4.1.4. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=125,OOO.

-112-
-c
p

_~~~.22'90 21.90
20.90
19.90
18.90
-c 17.90
p 17.40
16.90
16.40
15.90
14.90
13.9Q
12.90
11.90 .....,
9'.~090 e.e'?
8.90 cf
8.~090 LC,"-

-+.~090 0-r5
3.90 ,,"""
2.90 '<- (b)
, .90 rJ.
-o.~090 ,-e.
-2-." 0' 0 ,<-"0,
-3.10
-'i.l0
't-:-;fffl-+-~"-T-~---;"--t--~--"!;--~-""";r;' -5.' 0

-1

Figure 4.1.5. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=150,OOO

-113-
(\',---
(,
E_

"

-c

10

-c
~~~22'90 19.90
18.90
17.90
17 .... 0
21.90
20.90

p 16.90
16.-10
15.90
14,90
13.90
12.90
11.90 "]
91.~090 e.e'"
8.90
7.90
cl
Oe
S.90 L
'L~090 (Jr:F
3.90 . . .,)"""
2.90 ';-
1.90 J. (b)
-o.~090 '\..e
-2~'t 010 ,;-{'Co
-3.10
-".10
-5.10
9
Xl0- 1

-1

Figure 4.1.6. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=175,OOO.

-114-
5,

Ii
, I
I
!

3~

:: a)

22.90

1
21.90
20 .90
19.90
18.90
17.90
17.40
-c ! 16.90
p 4 16.40
r 15.90
15.'1"0
I 14.90
13.90
I
3'
12.80 1
1 \.90 eO
r 9',° 90
90 c§e
7.;090 L C?'
6.90 C
4 .~090 ,,"vCr
3.90 '(- (b)
2.90 0'
O.~090 "cz
_1-.~010 't'-.('~
-2.10
-3.10
-'4.10
-5.10
9
X10- 1

-1

Figure 4.1.7. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=200,OOO.

-115-
·I
I II
-c
p II

- \
'. .::1../

9 10
X10- 1

22.S0
21.90
20.90
19.90
18.90
18.'W
17.90
17."f0
16.90
-c 16.10
p 15.90
15.10
'1.90
I
14."10
13.90 e"
12'.;0"'0 C§e
10'.~090 L6Q.
9.90 C
7 .~090 '"vCr
Ei .90 ,,"
5 .90 J:
3 .~090 '-e
1 .~090 ,;-<'0,
0.90 (b)
-0.10
-1.10
-2.10
-3.10
-'L1D
-5.10
9
X10- 1

-I

Figure 4.1.8. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=2S0,OOO.

-116-
-c

22.90
21.90
20.90
!9.S0
18.90
18."\0
17.90
17 .'to
16.90
16.40
15.90
15."W
-c '" .90
H.40 I
p .,
13.90 eo'
13.40 (e
12.90 C;
12."W be,
11.90 1_

101.~O"lO (JC
10.'+0 ,,"'-'
9 .90 't-
8.90 J
S.~090 ",rL-

., .;090 >:-"0,
3.90 (b)
2.90
1.90
0.90
-0.10
-1.10
-2.10
-3.10
-4.10
-5.10

-I

Figure 4.1.9. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=300,OOO.

-117-
-5.10
-1.10
-3.10
-2.10
-1.1{)
-C.l0
0.90
1.90
2.90
3.90
-c 4.90
p 5.90
6.90
7.90 I
9 .~090 ri"'"
10.90 C§
1/~O'1-0 LeX"
12t.~O~O c:rC
13 .... 0 ","v
13.90 ';-
Hl.~"'O ox.
15 .... 0 '-c
'B'.~090 ,;-"c,
1 S .90
17.10
17.90
18.10
18.90
18.90
20.90
2T .90

22.90
21.90
20.90
IS.!'JO
113.'90
lB.10
17.90
17.'10
1B.90
16.'10
'5.90
15.40
-c Ii .90
p 11.40 '""r
13.90 e'"
'2'.~O"0 C§C
1t t.~010 ,-012
10'.~O"O (JC
10 ....0 "'-''''
9.90 't-
8.90 0'<
.~o90 ,-fl,
6
1.~o90 't-<,c, (b)
3.90
2.90
1.90
0.90
-0.10
-1.10
-2.10
-3.10
-4.10
h--Jftl......~--'\---7---.!,----T---,;----.!;--.:t;9:::=_Z -5.10
XlO- 1

-1

Figure 4.1.10. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=350,000.

-118-
-5.' C;

-c
p

9 ,0
Xl0- 1

22.90
21.90
20.90
19.90
19.10
18.90
18.10
17.90
17."f0
16.90
16.40
-c 15.90
p 15.40
1'4.90 ""I
131.~040 Q.e0
13.10 cf
90
'2'.;0 c.Oe
101.~090 U-
9.90 >vD-
e.90 't-......
6.~090 0"
5.90 ........e
3 .~090 't-"o, (b)
2 .90
1 .90
0.90
-0.10
-1.10
-2.10
-3.10
-1.10
-5.10

-1

Figure 4.1.11. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=400,OOO.

-119-
-c
p

( a)

-c
p

-,

Figure 4.1.12. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=450,OOO.

-120-
-5.10

I ~
:
=======---======::::;f -3.10
-~.10
?.:= / -0.10
_1-.~O'0
T 0.90
7 ~ .50
/ 2.90
3.90
"1.90
5.90
6.90
7.90 I
8.90 eO
9.~O~O de
-c
D 10'.~O"rO ,,-oe.
11.40
11.90 "C-
e-
12'.~"to ","
13.10 J:
H'.~090 ",Q,

151.~090 't-"c,
15.90
16.10 t .~)
lS.90
17.10
17.90
18 .... 0
18.90
'8.'tO
19.90
20 .... 0
20.90
21.90
22.90

ii'i-~ ~ ~ ~ ~=:~l ! 1)!;: t~;


21.90
20.90
20 ....0
19.90
19.10
18.90
18 ....0
17.90
17 .... 0
16.90
lS."rO 22.90
15.90
15.10 . ,
11.90 e'"
-c 131.~010 ~e
D
121.~010 L OIZ

11'.~O"'O Cr e
11 .... 0 "...J'><v
10 .. 90 't:-
10.10 J:
9.~090 \..(l
7 .~090 -.;-"c,
B.90
5.90
1.90
2.~090 (b)
1.90
0.90
-D.l0
-1.10
-2.10
-3.10
-1.10
-5.10

-1

Figure 4.1.13. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=500,OOO.

-121-
-c
p

-c
p

-1

Figure 4.1.14. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=550,OOO.

-122-
-c
p

s 19."0
19.90

~
~ 17.90
18."0
17."0
16.90
16.10
15.90
15 ..... 0
H.90
-c ,.... ..... 0
p 13.90
13.-10
12.90 1

1
11.<0
".~o·o cl e""
101.~090 L ~>",

9 .~090 "U
8.S0 ","
7 .90 "
5.~090 0'
..... 90 '-0 (b)
2 .~090 ,,('COl
1.90
0.90
-0.10
-1.10
-2.10
-3.10
-';"0
-S.10

-1

Figure 4.1.15. 3-D Plots of Cp vs x/c vs a for the Lower (a)


and Upper (b) Surface of a NACA-4415 Aerofoil
Section and Re=600,OOO.

-123-
;(10- 1
50

\ \
Angle of Attack

o : Uppe-r" SurFacE'
= 18'-:'.0 (dE-g.)

Q: Lo .... (>r SurFace

~ ,,

( 3.)

Angle oF' Atlack ;:: 18.~ Cde>g.,

0: Upper SurFace
Q: Lo ...... ,.. Sl.IrFace

~ ,,

(b)

Angle of I\t.t.ack = 18.90 (deg.;

o : Upppr SurFace
Q : Lo .... (>r SurFace

~ ,,

( C)

~igure 4.1.16. Pressure Distribution Around the NACA-441S


Aerofoil Section at Incidences Close to
Complete Stall (Re=300, 000) .

-124-
AngLe of ....t.t.ack '" 16.90 (deg.)

c : UPPE'!'" S .... rFacl'


<:) : Lo .... e!'" S .... rFO'cl'

] 3.)

1 "'ngle of ... Uack '" 17.10 (de-g.)

0: Upper SurFace
o : Lo .... er SurFace

-0
,

XIO-I

... ngte of Attack = 17.90 (de-g.)

0: Upper S .... rPacl'


Q: Lo .... er S....rFacli'

-0
,

(e)

Figure 4.1.17. Pressure Distribution Around the NACA-441S


Aerofoil Section at Incidences Close to
Complete Stall (Re=175,OOO).

-125-
~ngle oF' Attock = 18.90 (de-g.'

c : lJpper S .... rF acE'


<): lOIo'E"r SurFace

~, ~

~-~~~.-
XtO- t

1 Angle of Attack'" 19.10 <deg.)

0: Upper SurFace
<): Lower SurFace

~. (b)

~-~~~~
,:;====_l~_'_'
,~/~,c ~10-1

Angle oP ALtack '" 19.10 (O€'g.'


J.

o : UPPE"I'" SurPacE'
o : Lower SurFac@

-c ,

(C)

Figure 4.1.18. Pressure Distribution Around the NACA-4415


Aerofoil Section at Incidences Close to
Complete Stall (Re=400,OOO).

-126-
20 ,- ------------------------- ------~

ex 6
'-~
L

10
Q

~.,
f-'
l'J
---J
o-
-0- Lam i nar sep.
-+- Trans it i on
, ~,
.... Reattachment "tJ
~ Turbulent sep.

- I0 I --,-----.~~--~ -------,------~-

0.0 0.2 0.4 0.6 0.13 1.U


x/c

Figure 4.1.19. Locations of the Various Boundary Layer Phenomena on tile Upper
Surface of the NACA-4415 Aerofoil Section at Re=75,OOO.
20 -,-----

C<
l\\
~~~
10 .

f-'
tJ
CD o
-0- Laminarsep
-+- Transition ~"'., ""0-.-_._
... Reattachment
---0
-¢- Turbulent sep

-t---
- 10 --------,.----------.---------,------~-.----

0.0 0.2 0.4 U.6 u.8 1.0

x/c

Figure 4.1.20. Locations of the Various Boundary Layer Phenomena on the Urpel:
Surface of the NACA-441S Aerofoil Section at Re=100,OOO.
6

NACA-4415 AeroFoil
Angle of AUack = -5.10 (deg.)
5

-c
p

599790.3
550159.9
499570.4
450524.1
397925.5
348849.7
2880:51.1
248733.4
196919.7
173418.3
149128.7
124521 .3
99758.5
74077.0
19878.0

Figure 4.2.1. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
<X.=- 5.10°.
6

NACA-4415 AeroFoil
Angle of AUack = -3.10 (deg.)
5

-c
P 4

599702.8
550233.4
499328.9
450798.8
399086.9
349568.3
300'196.2
249704.4
198395.6
173621.1
149453.7
124742.9
99512.4
7'1206.3
50000.5
5 5 7 8

Figure 4.2.2. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
<X.=-3.100.

-129-
NACA-4415 A~roFoll

AngLe of Attack = -1.10 (deg.)

-c
p

599881.1
5502H.9
199800.1
150771.8
400050.5
349862.6
300023.9
250756.2
198748.2
171021.1
H9114.0
121608.4
100087.8
71272.7
50076.2

Figure 4.2.3. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-441S Aerofoil Section at
U=-1.10 o .
6

NACA-4415 A~roFoll
Angle of Attack = -0.10 (deg.)
5

-c
P 4

600362.3
550056.9
499356.7
452219.7
400509.9
350713.7
300488.4
251328.5
198968.4
171559.2
149537.0
124681 .9
100164.0
73936.6
50111.9
5 6 7 8 9
XlO- 1

-1

Figure 4.2.4. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-441S Aerofoil Section at
U=-0.10 o .

-130-
6

NACA-4415 AeroFolL
5 AngLe of Altack = 0.90 (deg.)

-c
p

60030"1.1
550622.0
2 500030.2
151S57.0
"100"133.9
350421 .5
301009.3
251887.4
199370.4
174385.6
1"19595.6
12"1731.3
100231.7
73999.7
7
50134.8
8

-1

Figure 4.2.5. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
U= 0.90 0 .
6

NACA-4415 AeroFoiL
AngLe of Altack = 2.90 (deg.)
5

-c
P 4

600818.3
551160.4
2
501079.8
151924.8
400987.1
3532H.l
300029.3
25189"1.8
198400.6
174735.4
H9963.3
124513.6
100366.4
74140.8
50170.4

-1

Figure 4.2.6. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-44l5 Aerofoil Section at
U= 2.90 0 .

-131-
6

NACA-4415 AeroFolL
AngLe of Attack = 4.90 (deg.)
5

-r:
p 4

601826.7
551709.9
500469.9
451479.5
400904.0
351788.2
300758.2
252148.4
199082.6
174852.8
150323.3
124597.2
100093.1
74810.0
o 50206.0

-1

Figure 4.2.7. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
a= 4.90°.

NACA-4415 AeroFoiL
Angle of Attack = 6.90 (deg.)

-c
p

601315.6
551877.1
500540.3
451318.7
401186.1
352218.2
301310.9
252249.2
199560.1
175099.8
150520.1
124975.3
100193.5
74946.4
0~---t----22----~3----~4----~5;---~6~--'7~--~8~---9~--~ 50213 .2
XIO-I

-1

Figure 4.2.8. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-441S Aerofoil Section at
a=6. 90°.

-132-
6

NACA-44J5 AeroFoll
AngLe of Attack = 9.90 (deg.)
5

-c
P "I

-1

Figure 4.2.9. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
U= 9.90°.
6

NACA-4415 AeroFolL
5 AngLe oF' Attack = 10.90 (deg.)

-c
P "I

6017"13.3
552031.7
2 501 32·LI
"151108.2
"101858.1
353228.2
301109.3
252680.7
199836.6
174967.1
150637.5
125082.0
100302.8
7+583.0
°t----,----?2----~3----~4----~5----~6~--~7,---'8~--'9~--~ 50293.1
XlO- 1

-1

Figure 4.2.10. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
U=10. 90°.

-133-
6

NACA-4415 AeroFolL
AngLe of Attack = I I .90 'Geg. )
5

-c
p

5 5 7 8

-I

Figure 4.2.11. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-441S Aerofoil Section at
Ct=11.900.

NACA-4415 AeroFolL

5
AngLe of Attack = 12.90 (deg.)

-c
P 1

601331.7
551216.1
2 500982.3
151317.4
401786.8
352835.3
300374.8
252+33.3
199583.1
175008.2
150451.1
125128.1
100408.0
7+675.5
o 50255.6
2 3 5 6 7 8 9
XIO-1

-I

Figure 4.2.12. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-441S Aerofoil Section at
Ct=12. 90°.

-134-
6

NACA-1415 AeroPolL
5 AngLe oF' Allack = 13.'30 (deg.)

-c
P 4

~~
~-------~---
3

~~ 50~~~~8.~.7
600888.6
2 l}
------- 1
_
--------
.~
/
/ 151304.4
400816.3
352880.1
---_____ 300651.2
253017.2
199171.8
175022.8
150179.0
12-1801.8
100473.4
74640.2
O~--t--22--~3--~4C-_-~5;--~6~-~7~-~8~-~9--}10 50240.3
XIO-1

-1

Figure 4.2.13. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
<X=13. 90 0 .

NACA-1415 AerOPOlL

5
AnqLe oF' Attack = 11.'30 (deg.)

-c
P 1

601181.8
552293.2
2 500383.6
151254.5
100399.1
352136.2
300350.8
252135.6
198961 .2
174763.6
150196.9
121801 .8
100540.6
71710.4
o 50229.9
2 3 5 6 7 8

-1

Figure 4.2.14. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
<X=14. 90 0 .

-135-
NACA-4415 AeroFolL

5 AngLe of Altack = 15.90 (deg.)

-c
P "I

~
~
3

601363.7
552223."1
500292.7
"\50344.0
"100189.6
351369.5
298836.8
251183.2
198509.0
174518.7
150507.5
12517"1.1
100237.8
7"1714.1
0~---t----~2----13----~4~'--~5;---~6~--~7~--~8~--~9----~ 50241 .2
XlO- 1

-1

Figure 4.2.15. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
U=15. 90 0 .

NACA-4415 AeroFolL
AngLe of Altack = 16.90 (deg.)
5

-c
P 4

601336.3
5518"i1.0
"i99796.7
2 45012"i.6
400047.1
351260.1
299543.1
250208.0
197425.4
174320.3
150523.4
124856.1
98718.0
73585.0
50226.5
2 3 4 ·5 6 7 8 9
Xl0-l

x/c
-I

Figure 4.2.16. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
U=16. 90 0 .

-136-
6

NACA-4415 AeroFolL
5 AngLe of Attack = 17.90 (deg.)

-c
p

599911 .~
550945.8
499589.8
4,9139.7
399575.3
350742.7
239163.6
249041.3
194120.8
172843.1
149477.5
122195.6
97890.2
71119.7
0~---T----~2----~3----~4----~5----~6~--'7~--~8~---9~--~ 50227.2
XIO-I

-I

Figure 4.2.17. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
Cl=17.90 0 .

NACA-4415 AeroFolL
AngLe oF' Allack = 18.90 (deg.)
5

-c
P 4

600365.3
550603. ,
498371 .1
39W>;8.~.0 e'
343537.5 ~
291719.8 ~0
243017.3
193079.3 00
172561 .3 '--
149747.6 ~o
122141.8 e.-0
98291.0 q.:
73562.8
0~--·-L----~2----~3----74----~5----~6~--~7~--·8~---'9~---fIO 50222.0
XIO-I

-I

Figure 4.2.18. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
Cl=18. 90 0 .

-137-
6

NACA-4415 AeroFoll
AngLe of Attack = 19.90 .: ,jeg. )
5

-c
p

496766.4
437484.3
390825.2
1." 24~~ ~~5.~.5
343014.4
lH::::::::::=:======================:::::
~ /
. 192032.6
172054.5
14;;503.1
122121.4
,,0
.-0
~
00

98281. 1 ,~I.&
74102.3
o 50218.6
2 3 4 5 6 7 8 9 10
XIO- 1

-1

Figure 4.2.19. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
0:=19.90 0 .

NACA-4415 AeroFoil
Angle of Attack = 20.90 (deg.)

-c
p

497247.5
'147306.1
390841.9
3'12204.0
291077.7
2'11852.7
191208.9
171568.9
150301.'1
121766.9
97887.2
74089.8
0~--~--~----~3----·t---~5----~6----~7~--~8'---~9~--~ 50214.3
XIO- 1

-1

Figure 4.2.20. 3-D Plot of Cp vs x/c vs Re for the Upper


Surface of the NACA-4415 Aerofoil Section at
0:=20.90 0 .

-138-
Minimum
pressure
peak~

~ \
I ,
I ',T'
.... -- ..... T
I
I f \r t(
Actual. pressure
vanOTlon with
I S \ separafJon
-cp I \ bubbfe
( \
I \
I T"lrbulent boundary \
I loyer pressure
I
'yariatlon without
seporotlcn bubble R
( trip wIre)

POSITION ALONG CHORD, X/C

Figure 4.2.21. Typical Surface Pressure Distributions With


and Without Separation Bubble.
(Adapted from Reference 12)

- 139 -
( a) (b) (c) (d) (e)

I
.........
~
o
I

Re=200,OOO Re=300,OOO Re=400,OOO Re=500,OOO Re=600,OOO

Figure 4.3.1. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at U=-5.10 o .
(a) (b) (c) ( d) ( e)
.,

.+:0.
..........
I

Re=2S0,OOO Re=350,OOO Re=450,OOO Re=500,OOO Re=600,OOO

Figure 4.3.2. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at U=-1.10 o .
(b) (c) ( d) ( e)
(a)

.........
~
N
I

Re=200,OOO Re=300,OOO Re=400,OOO Re=500,OOO Re=600,OOO

Figure 4.3.3. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at U=2.90o.
( a) (b) (c) ( d) ( e)

- I
..j:;:..
w
I

Re=200,OOO Re=300,OOO Re=400,OOO Re=500,OOO Re=600,OOO

Figure 4.3.4. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at U=6.90o.
(a) (b) (c) ( d) ( e)

-
I
~
~
I

Re=200,OOO Re=300,OOO Re:::::400,OOO Re=4S0,OOO Re=SOO,OOO

Figure 4.3.5. Flow Visualisation Photographs of the Upper Surface of the NACA-441S
Aerofoil Section at Various Reynolds Numbers and at U:::::9.90o.
..p..
Vl
I

( a) (b) (c)

Figure 4.3.6. A Sequence of Flow Visualisation Photographs of the Upper Surface of the
NACA-4415 Aerofoil Section at Re=600,OOO and at a=9.90o.
( a) (b) (c) ( d) ( e)

+:-.
0'1
I

Re=250,OOO Re=300,OOO Re=400,OOO Re=450,OOO Re=500,OOO

Figure 4.3.7. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at a=11.90o.
I
...-
+:-.
-...l
I

( a) (b) (c)

Figure 4.3.8. A Sequence of Flow Visualisation Photographs of the Upper Surface of the
NACA-4415 Aerofoil Section at Re=600,OOO and at U=11.90o.
( a) (b) (c) ( d) ( e)

~
00
I

Re=250,OOO Re=300,OOO Re=400,OOO Re=500,OOO Re=600,OOO

Figure 4.3.9. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at a=13.90 o .
(a) (b) (c) ( d) ( e)

~
\.0
I

Re==250,OOO Re==350,OOO Re==450,OOO Re==500,OOO Re==600,OOO

Figure 4.3.10. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at a==15.90 o .
(a) (b) (c) ( d) (e)

Vl
o
I

Re=250,OOO Re=300,OOO Re=400,OOO Re=500,OOO Re=600,OOO

Figure 4.3.11. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at a=17.90o.
( a) (b) (c) ( d) ( e)

--
VI

Re=300,OOO Re=400,OOO Re=450,OOO Re=500,OOO Re=600,OOO

Figure 4.3.12. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil Section at Various Reynolds Numbers and at a=18.909.
(a) (b) (c) ( d) (e)

-
I
VI
tv
I

Re=300,OOO Re=350,OOO Re=400,OOO Re=450,OOO Re=600,OOO

Figure 4.3.13. Flow Visualisation Photographs of the Upper Surface of the NACA-4415
Aerofoil section at Various Reynolds Numbers and at U=19.90o.
V1
w
I

( a) (b) ( c) ( d)

Figure 4.3.14. A Sequence of Flow Visualisation Photographs of the Upper Surface of the
NACA-4415 Aerofoil Section at Re=450,OOO and at U=19,90o,
20

ex
t
~~.
10
"' ~~~~~~\ ~
f-'
U1
o ~
~~"'- ~
.1'>

~*,'
-0- Larninarsep } o
-+- Reattachment Cp Data
.... Turbulent sep ~ ""'~
~ Laminarsep } ~ "",-

--- Reattacliment Flow Vis. Data '''''9


-0- Turbulent sep.

-1 0 ~t~~-----,- r"~'-~---r------''----'-'-'-'-T -.-.--.... ~-- . - . , ..-.--._.-. - .. -. ----,


0.0 0.2 0.4 CI.b 0.8 ] .0
z/c

Figure 4.3.15. Comparison Between Cp and Flow Visualisation Data Regarding the
Locations of the Various Boundary Layer Phenomena Occurring on
the Upper Surface of the NACA~4415 Model at Re=lSO,OOO.
20 -

IX

10-

"~
~

J-'
Ul
Ul o
~ ~"
-{r Laminarsep. } '~
+ Reattachment Cp Data ~
'--'--..
-. Tur·buientsep.
-<>- Laminar sep. } "-.-........,.~~. '-9
-II- Reattacilment Flow Vis. Data
-0- Turbu 1ent sep.

--.----- -.--. -,. ,_.. - -- - --. -- - - I


-10+- 1-
~-----,-------.-

0.0 0.2 0.4 0.6 o.e 1.0


x/c

Figure 4.3.16. Comparison Between Cp and Flow Visualisation Data Regarding the
Locations of the Various Boundary Layer Phenomena Occurring on
the Upper Surface of the NACA-4415 Model at Re=200,OOO.
20 -

ex

10

t-->
U1
0\
o
-rt
+
Laminar sep
Reattachment
}-c p Data --"
... Turbulentsep. ~"
~""""-"""'-"--""
-<>- Laminarsep. }
. . Reattachment Flow Vis_ Data ~~~
-0- Turbu 1ent sep.

- 10 +-------T 1 -------,-- 1 - ---- --,--- - - ..---------y---- ----- - -1'- -I

0.0 O.~~ 0.4 U.6 0.8 1.0


x/c

Figure 4.3.17. Comparison Between Cp and Flovv Vi~311alisation Data Regarclj ng th<:j
Locations of the Various Boundary Layer Phenomena Occurring ell

the Upper Surface of the NACA-4415 Model at Re=250,OOO.


IX

10

f-'
(Jl
--J
o ~¢- Lam i nar sep. }
+ Reattachment Cp Data
-+- Turbulent sep
+ Laminarsep. }
.... Reattachment Flow Vis. Data 1:1I
-{]- Turbu 1ent sep.

- 10 +-.~~-.-.. T--·-~~--'------ r- -- - ___ 0. _ _ _ _ _ _ T - .--- - -- ---~ --I~-

nn 0.2 0.4 0.6 0.8 1.0


x/c

Figure 4.3.18. Comparison Between Cp and Flow Visualisation Data Regarding ttlU
Locations of the Various Boundary Layer Phenomena Occurring on
the Upper Surface of the NACA~44]5 Model at Re=300,OOO.
20

ex:

10-

f-'
Ul

o-
¢-
+
Laminar sep. } 1
\
Reattachment Cp Data
... Turbulent sep
-¢- Laminarsep. } -""'-' \
--It- Reattacliment Flow Vis. Data "¢ ~
-{]- Tur'bu lent sep

----- -1- --_. ----._- _.- - -- ----,-.---- -------1


-10 --~----"-,... -r-------~ -,.......---~-- -T- ------

0.0 0.2 0.4 0.6 0.8 1 . \)


x/c

Figure 4.3.19. Comparison Between Cp and Flo\', Visualisation Data RegauHn<;;r Llil~

Loca tions of the Various Boundal:y Layer Phenomena OccurI.' in9 ,)n
the Upper Surface of the NACA-4415 Model at Re=350,OOO.
20 - _1

ex

10-

f-' "

lJl
o ".
,
~
"

\D
LarninJrSe p. } "1A ,
+ Reattachrnent Cp Data

-¢- t
. . Turbulent sep.
Larninarsep.
-II- Reattachrnent
-{]- Turbul ent sep
Flow Vis. Data
\
\

~J

-------------,--~--·-----t--~-------~ --1"- - ------ - -l- -- - - - - - - - - -.. - - , - - ------- - -,- - - - - ---1


-I01---~

0.0 0.2 0.4 o . c' 0.[\ 1.0


xl c:

Figure 4.3.20. Comparison Between Cp and Flow Vi sual:Lsa Lion Data Regardi li<;J the
Locations of the Various Boundary Layer Phenomena Occunin<;'1 on
the Upper Surface of the NACA-4415 Nodel at Re=400,OOO.
20

C<

"~"
10- ~¢L"",
"'~

'~

I--'
y~
Q')
o
o- -¢- Laminarsep }
+ Reattachment cp Data "~
... Turbulent sep. "~
-<>- Laminarsep } ",
.. Reattachment Flow Vis. Data '-'''''9
-0 Turbulent sep

- 10 -f-~~------.----- -'--~'----T y---~~~---- ------ - r------ --~-- ---l

0.0 0.2 0.4 0.6 () • 13 1.0


x/c

Figure 4.3.21. Comparison Between Cp and Flow Vi sualisaU on DaLa Peg,ell di Ij';; til
Locations of the Various Boundary Layer Phenomena OccuJ:rirl9 on
the Upper Surface of the NACA-4415 Model at Re=500,OOO.
20·

(X

10

f-'
O'l
f-'
o -¢- Lam i nar sep. }
+ Reattachment Cp Data
-t Turbu 1ent sep. \\
-<>- Laminar'sep, } \
-III- Reattachment Flow Vis, Data 't;J
-0- Turbulent sep,

-1 0 -I--~---~~-. -----,-------,--.- -- .. -~- ~- -_.,.-


0.0 0.2 0.4 0.6 0.8 1.0
x/c

Figure 4.3.22. Comparison Between Cp and Flow Visualisation Data Regarding tile
Locations of the Various Boundary Layer Phenomena Occurring on
the Upper Surface of the NACA-4415 Modol at Re=600,OOO.
X10-1
14

12

10

C 8
n

REYNOLDS NUMBER = 50000.0


6
¢ : UNCORRECTEO O~T~

AngLe (Degrees)

-10 -5 5 10 15 20 25

J
AngLe (Degrees)

-10 -5 5 10 15 20 25 3

-5

-15

XlO- 1
3

-c t
2

AngLe (Degrees)

~100--------~~~_-------~~0----~~=:J~t=~,~.~,-:,=~A~~~~~~~~~~t~[;.~,~~~.~'~£~G~~.~'=_~--;2~5;-------~3C
~&- () -+~-

Figure 4.4.1. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=50,OOO.

-162-
X10-1
14

12

10

C 8
n

REYNOLDS NUMBER = 75000.0


6
.,: UNCORRECTED DATA

Angle (Degrees)

-10 -5 5 10 15 20 25 30

-2

Angle (Degrees)

':'10 5 10 15 20 25 30

-5

-15

XlO- 1
3

-ct
2

Angle (Degrees)

-~1~0~------~=-------~~~------5S---------~1~O~------~155--i:~:;~~=:==~~2~5r--------:30

Figure 4.4.2. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=75,OOO.

-163-
X10-1
14

12

10

C 8
n

REYNOLDS NUMBER = 100000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 -5 5 10 15 20 25 30

-2

Angle (Degrees)

-~1nO~------~-~5--------~---------'5t---------'1~O~------~;~5--------~2AO--------~2~[5~------~30

-5

c
m
1/4
-15

-c t
2

Angle (Degrees)

Figure 4.4.3. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=100,OOO.

-164-
12

10

C 8
n

REYNOLDS NUMBER = 125000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 -5 5 10 15 20 25 30

-2

Angle (Degrees)

-~11~0~------~~~5--------~----------~5--------~1~O'-------~1~5--------~2~O-------~--------~30

-5

-15

-c t
2

Angle (Degrees)

-~1~0--------~-~5--------~~~------~5--------~1~O'-------~1~5----~--~~--------~25~------~30

Figure 4.4.4. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=125,OOO.

-165-
XlO- 1
14

12

10

C 8
n

REYNOLDS NUMBER = 150000.0


6
0: UNCORRECTED OATA

AngLe (Degrees)

-10 5 10 15 20 25

-2

Ang L e ( Degrees)

-10 -5 5 10 15 20 25 3

-5

-15

XlO- 1
3

-C
t
2

Ang Le ( Degrees)

-10 -5 5 10 15 25 3D

Figure 4.4.5. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=150,OOO.

-166-
12

10

C 8
n

REYNOLDS NUMBER = 175000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 5 10 15 20 25 30

-2

Angle (Degrees)

-'"10 -'"5 5 10 15 20 25 30

-5

m
1/4
-15

XlO- 1
3

-c t
2

Angle (Degrees)

-10 5 10 15 25 30

Figure 4.4.6. Normal Force, Quarter Chord pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=175,OOO.

-167-
Xl0-l
14

12

10

C 8
n

REYNOLDS NUMBER = 200000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 5 10 15 20 25 30

-2

Angle (Degrees)

-10 -5 --~5---------~i~0--------'i~5~------~2~[O--------~25~------~30

-5

XlO- 1
3

-c t
2

Angle (Degrees)

-10 5 10 15 o 25 30

Figure 4.4.7. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=200,OOO.

-168-
X10-1
14

12

10

C 8
n

REYNOLDS NUMBER = 250000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 5 10 15 20 25 3

Angle (Degrees)

-~100------~-~5'-------~--------~5;-------'~1~fO;--------ti;5~------~2~0--------~2~5--------~3C

-5

-15

-ct
2

Angle (Degrees)

-10 5 10 15 o 25 3D

Figure 4.4.8. Normal Force, Quarter Chord pitching Moment


and Tangential Force Coefficient Variation
with Angle of Attack at Re=2S0,OOO.

-169-
X10-1
14

12

10

C 8
n

REYNOLDS NUMBER = 300000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 5 10 15 20 25 30

-2

Angle (Degrees)
X10- 2
-~1~0--------~-5~------~--------~5--------~1~O~-------~i5~------~2~0--------~2~5--------~30

-5

-15

X10- 1
3

-c
t
2

Angle (Degrees)

-10 5 10 15 o 25 30

Figure 4.4.9. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=300,000.

-170-
Xl0-l
14

12

10

C 8
n

REYNOLDS NUMBER = 350DOO.0


6
0: UNCORRECTED DATA

Angle (Degrees)
I I I I
-10 -5 5 10 15 20 25 3(

J
Angle (Degrees)

-~1~0---------~5~------~----·----~5~------~lh'O~------~;5~------~2AO--------~2~5~------~3l

-5

-15

XlO- 1 .
3

-C
t
2

Angle (Degrees)

-~100------~~;=~;=~~~------~5'---------lho~-------t15~-----t:j~O=====---;2~5'--------:30

Figure 4.4.10. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=350,OOO.

-171-
XlO- 1
14

12

10

C 8
n

REYNOLDS NUMBER = 400000.0


6
~: UNCORRECTEO DATA

Angle (Degrees)

-10 5 10 15 20 25 3C

Angle (Degrees)

:w-------~-~5--------~---------.5~--------~Ino---------+i5~------~2~10,-------~2~5~--------3o

-5

c
m
1/4
-15

XlO- 1
3

-ct
2

Angle (Degrees)

-10 5 10 15 o 25 30

Figure 4.4.11. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=400,OOO.

-172-
X10-1
14

12

10

C 8
n

REYNOLDS NUMBER = 450000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-~1~O--------~-5~------~--------~5~--------~10'--------+15~------~2~O--------~2~!5~------~30

-2

Angle (Degrees)

-~1~0~------~-5~------~~-------O5~!---------~io,-------·-+i~5--------~2~O--------~2b!5~------~30

-5

-10
c

-15

-c
t
2

Angle (Degrees)

-~1~O------~~~~;=~t-~------~5s---------nl0~-------t15~-------i~------~2~5;-------~30

Figure 4.4.12. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=450,OOO.

-173-
XIO-I
14

12

10

C 8
n

REYNOLDS NUMBER = 500000.0


6
~: UNCORRECTED DATA

AngLe (Degrees)

-10 -5 5 10 15 20 25 30

-2

AngLe (Degrees)

-~1~0--------~-5r-------~r----------'5~'--------~i~0--------~i5r---------2~'O--------~2~5---------30

-5

-10
c

-15

XIO- 1
3

- C
t
2

AngLe (Degrees)

-~IOO--------~~~~~~~~-----'5S---------~100---------t155---------22~t:,::,==.---:2~5~-------:30

Figure 4.4.13. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=500,OOO.

-174-
12

10

C 8
n

REYNOLDS NUMBER = 550000.0


6
0: UNCORRECTED DATA

Angle (Degrees)
I I
-10 -5 5 10 15 20 25 ~

-1
Angle (Degrees)

~-10--------~-5~------~--------~5~--------~10'-------~15~------~2~O--------~2~5~------~3C

-5

-c
t
2

Angle (Degrees)

Figure 4.4.14. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=550,OOO.

-175-
12

10

C 8
n

REYNOLDS NUMBER = 600000.0


6
0: UNCORRECTED DATA

Angle (Degrees)

-10 5 10 15 20 25 30

Angle (Degrees)

-10 -5 5 10 15 20 25 30

-5

-1

c
m
1/4
-15

XlO- 1
3

-C
t.
2

Angle (Degrees)

-~100--------~;=:=~~~~------~55---------~10~-------t155--------22~0--------:2~5;--------:30

Figure 4.4.15. Normal Force, Quarter Chord Pitching Moment


and Tangential Force Coefficient Variation
With Angle of Attack at Re=600,OOO.

-176-
Wortmann · 3T C X10- 1
FX 63-137 mc/4 1~ CL,2oo V~ ALPHA
Corrected data RETNOLOS NO. 320225.
Rc = 99, 388 CL
c = 6 in. · 2-
Smooth 2CC

· 1

Alpha, degre8s
"---r
1_
~,----- .. --
__J

4 12 20 28

I---' 200
--.J
--.J

-10 10 15 20 25 30

ANGl.£ OF ATTACK (deree.o I


-.3

Fig. 4.4.16. Qurter Chord Pitching Moment Coef- Fig. 4.4.17. NACA-23012 Aerofoil Data for
icient Versus Angle of Attack for AR=4.0.
the Wortmann FX-63-137 Aerofoil
(Rc=100,OOO) (Adapted from Reference 36)

(Adapted from Reference 26)


10S'r----r----.-----
c,
1.2

10
'1 J~

0.8

Cl

0.8 .~ JP

Geometric AR 2.466
f-' ........... Geometric AR 2.0
--J . _ . _ . Geometric AR 1.5
CD o
- - Geometric AR 1.0
- - - - - Geometric AR 0.5 "'I II
5
RC·ZoO·'0
- 0.4

-10 o 10 20 ____________
30
-I... ...
-.S'~ L__ _ _ _ _ _ _ _ _ _ _ _ ~ ____________ ~

()(,o I I•• 2 •. '


a (dig.)

Fig. 4.4.18. Lift Coefficient Versus Uncorrected Fig. 4.4.19. Measured Aerodynamic Characteri-
Incidence of a NACA-23012 Aerofoil stics of a NACA-23012 Airfoil.
at a Reynolds Number of 350,000.
(Adapted from Reference 38)
(Adapted from Reference 37)
1· 5

1· 4 I
,.. ,
'- ""

0 - Lower Surface Exhibits CNmax .


Fully Laminar Attached 1. 3
Flow
SOl x - Occurrence of "kink"
ex 6° 1·2
I-'
--J

\.0
0 Q Q Q ~
2° X S? ~ Q Q 8

2 3 4 5 6
- 2° -s
Re x 10
- 4°

0'°0 2 3 4 5 6

5
Re x 10-

Fig. 4.4.20. Comparison of Results for the Occur- Fi<;r. 4.'5.1. [\']ct);irnum NCILIII.11 ]'(1)(,'", (_>j,dijc'ic'?lll

rence of the 'kink' on C and C Var ia tjOll VJi th Rt'~'nolc:L; Number.


N Mc/4
Curves With Those Defining Fully
Laminar Flow Using Caton's Method.
.5
Wortmann
FX 63-137
Rc"77.6D4 1.6
Uncorrgctgd Data I /\ \ . 4
Cd
C 1. 2 '

1 I J ~ . •3

.8
I .-1> ~
.2

.. :Ii A' , --'---'


20
~,-r~
-20 4 12 28
Al pho. tJQ9rQI;H~ ... ---+--
-20 -12 -4 4 12 20 23
-. 4
" 1pho. d;;;>groQs

Figure 4.4.21. Two Dimensional Lift and Drag Coefficients versus Angle of
Attack, Re = 80,000.

(Adapted form Reference 33)


1 ·4

1 ·2

1 ·0

CL max REYNOLDS No
1 28 700,000 •
7 1,000,000
0-4 0

+
i 1· 42
1· 46
1,500,000
2,000,000 x

10'2 1 ·48 3,000,000 ..

g-4 4 8 20 24
.-i
1'2 16

-0'2

~
• iii
-0,4

Figure 4.5.2 Lift Coefficient Variation with Incidence and


Reynolds Number of a NACA-4415 Aerofoil
Tested at Stuttgart (1962-72).

(Adapted from Reference 29)

-180-
J~---~----~---20
5to. i Up'r. i L'wi-: , y h~ .l a .c . 1
1 1
l' /0 J:lli ~
l/ r 1'--1-+--
0~
1 ;::
O. - I 0 '- , ! I
1.25; 3.07; -1.79 i ~.g .... 't--.

2.51 4. /7 -2.48 I' ; . , .... \.J -10 !, II ',J I i I 1

5.015.74 -3.27 ~ " :', )1


7.51 6.9/ - 3. 7 / i 0:: 0 I ~, I : I 1"'cf41 i I

/017.84 -3.98 I 0 20 40 60 80 100


,
/5 9.27 -4. /8 I I Fercenf ofchord
20110.25 -4. /5
25 1092 -3.98 - - - r - - - - - - , - - ,;---'-;-~-----
I f-i

i : i
30:1 1.25! -3.75 : i I I

401//. 25 3 .25 I !
50:10.53 -272
1-
i
,
I
I

i :
I
1 i i !
,
!
2.0 t
t

60 9.30 i -2.14 I i i ! i i I I i I
70 7.63 i -/.55 i i ,
I i i I ! I 1.8
80 5.55 -1.03
I I : I
I
,
I 1

I
9013.08 - .57 j I I
951/.67 - .36 ~ ..tY::.
1.6
! IDa i (.16) (-.16) :
o I
I I
I
t d ~. rn~
I

"
::s,.(- 1

- to ~-~-x, r--A I
I
:
11001 , :
I
!
~- K
I :

! L E. Rod.: 2.48 : J
I
I

~ 'L~ '- ~
7'
... 1 I
, t
,
I
I
1.4
, Slope of radiusi
tl"lrough end of! ! t1 v/......
'~'L~"
....,.. , "
~ ...
'-J,
;"'1
...
V I
i I
1.2 t.J
chord: 4/20 I ' A~V "'" ~"- ~
I

~ ~'
fo II I
'\
" >, ru.
~ ~" ''\
''tz
I.,
'"I,

r-.....
" ........
I-l- !-
~
I
;.,.-
c:::
1.0 ......
Q;

~ ~ fo. ...'J' .. - :-- -= .ll


~
8 ~Q)
,...,

Test ~v
I---
j/ •
o
.1'
,., Reynolds Number_
.6 u
~
0 3,000,000 .....
\t:
6.-----2,380,000
x----------- 1.260,000 .4-.....J
IYi" + - - - - - - 654,000
1P \7------- 334,000
0------ 163,000 .2
1
'{,'I
1/
'\J--------- 83,000
t £7--------41 700 o
J~ Airfoil: NA.C.A. 44/5

Ifl /
i~,

~
" Size: 5"x30"

Test: V. D. T / /63
Vel. (ff./sec.):69 -.2
Rres.(sfhd. afm.): 1/4 fo 20

Where fesfed: L.MA.L.


Dafe: 8-34
-,4

-8 -4 0 4 8 /2 16 20 24 28 32
.Angle of attack for infinife aspecf ratio, eto (degrees)

Figure 4.5.3 Lift Coefficient Variation with Incidence and


Reynolds Number of a NACA-4415 Aerofoil
Tested at NACA VDT (1934).

(Adapted from Reference 34)

-181-
1· 4

1 ·2

1·0

0·8

(Lmax REYNOLDS NO.

1 -38 0 700,000
1·36
1·37
• 1,000,000
0 1,500,00 °
1·44 x 2,000,000
1·34- + 3,000,000

4 8 12 16 20
a

- 0· 2
II

x+
+ -0· 4

Figure 4.5.4 Lift Coefficient Variation with Incidence and


Reynolds Number of a NACA-4415 Aerofoil
Tested at NACA LTT (1945).

(Adapted from Reference 28)

-182-
X10-1
16~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~

14

12

ReynoLds Number
4 II : 150,000
v: 175,000
+ : 200,000

2
x : 250,000
0: 300,000
<> : 350,000

°
NACA-1~15 AeroFoil
-2
UNCORRECTED DATA

-4

( a)

Figure 4.5.5 Normal Force Coefficient Variation with


Incidence and Reynolds Number of a NACA-4415
Aerofoil Tested at University of Glasgow
(1987) .

-183-
XlO- 1
18~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~

14

12

Reynolds Number
4 ,,: 400,000
v: 450,000
+ : 500,000
x: 550,000
2
0: 600,000

o+-__ ______-+________________________________________________
~ ~

NACA-~~15 AeroFoil
-2
UNCORRECTED DATA

-4

a
(b)

Figure 4.5.5. Concluded

-184-
Wortmann 1.6
FX 63-137
CcrrGctad data
Rc = 199. 487
c lC6 1 n. 1.2
Smooth

-20
;::J-4 4 12
Alpha.
20
dQgrQQs
28

-. 4

Figure 4.5.6 Lift Coefficient Variation with Incidence of a


Wortmann FX-63-137 Aerofoil Tested at a
Reynolds Number of 200,000.

(Adapted from Reference 26)

-185-
NACA
1· 4 LTT
'x

1·2
(L GLASGOW

1· 0

0·8

TEST CONDITIONS
TUNNEL NACA LTT STUlTGARHI1 GLASGOW
TEST 2-D 2-D 2-D
0·03% 0·02% ~O· 5%
3 ·81 5'46 2·80
70~OOO 600,000
SMOOTH SMOOTH

-8 4 8 12 16 20 24

a. ( 0)
,. - 0·2
~.

If
- 0'4
a. ( 0)
(MC/I.
-8 -4 4 8 12 16 20 24

-0'04

..
-x-X _x-x-.x-;X~.··x
.. ' .• . x 'X

•...•...•.. -0·12 •••


.. ' ....•...•... •••
'

Figure 4.6. Comparison of Lift and Quarter Chord Pitching


Moment Coefficients of the NACA-4415 Aerofoil
Obtained at Different Test Environments and Wind
Tunnels.

-186-
'2.2 CN CONTOURS

'·0 1-0
20
a (0)

18
1-20

16 1-23
1-25
1-28
14 1-30
1·31
1-32
-12
1-32
1-31
1'30
10 -28
1-25
1-23
1·20
8 1·18
1-15
1·10
6
1-0

0-9
4
0-8

0-7
2
0-6
0-5
6
·4
0-3
-2 0·2
~0'1
-4 0-0

-5
- 0'1
Re x 10

Figure 4.7.1. Normal Force Coefficient Contours for the


NACA-4415 Aerofoil with Incidence and Reynolds
Number.

-187-
~ '?
24 <:)

1- 5

22

20

0-8
18

0-7

16

14 0-6
0<.0

12
0-5
~'6
10 0-4

8
0-3
1-4
6
0-2

4 0-'

0-0 '·0
2
-----0·9
0-8
-0-'
0-7
0 6
0-6

-2 o·s
0-4

Figure 4.7.2. Normal Force Coefficient Contours for the


GU25-5(11)8 Aerofoil with Incidence and
Reynolds Number.

(Adapted from Reference 35)

-188-
18
0·97
1·06
1 ·17
16

14

1·17
12

10
1· 06-

8
0·93

6
0·81

4 0·70

0·58
2
0·47

0 0·34
3 5
0·23
-2
0·11
RE x 10- 5

-4

Figure 4.7.3. Normal Force Coefficient Contours for the


NASA GA(W) -1 Aerofoil with Incidence and
Reynolds Number.

(Adapted from Reference 18)

-189-
16
0'6
10'8
~0'7
.... O·g
14

o·g

0·8

_ 0'7
~
~

~0'6
..............
. / ' 0'5
----- 0'4

2
----
~
- 0'2
0'3

0·1

1 2 3 5

Re x 10- 5

Figure 4.7.4. Normal Force Coefficient Contours for the


NACA-0015 Aerofoil with Incidence and Reynolds
Number.

(Adapted from Reference 18)

-190-
EPo~Y Resin foam

Gel-,~oat surface
Aluminium End Plate

Gla S \3 fibre wa~ Mould

Al\1lnininm End plate

B3700d
S
Locating p}n

I-'
\.0
I-'
I

~1\O.· ~ 41;)
et
. f the Mod

Figure bl. constcuctlOn 0

You might also like