Vous êtes sur la page 1sur 165

THREE DIMENSIONAL AERODYNAMICS OF A SIMPLE WING IN OSCILLATION

INCLUDING EFFECTS OF VORTEX GENERATORS

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree of Doctor of Philosophy in the

Graduate School of the Ohio State University

by

Jolanta M. Janiszewska, B.S.A.A.E., M.S.

*****

The Ohio State University

2004

Dissertation Committee: Approved by

Dr. Gerald Gregorek, Adviser

Dr. John Lee ____________________________


Adviser
Dr. Richard Bodonyi Aeronautical and Astronautical Engineering

Dr. James Scott Graduate Program


ABSTRACT

A comprehensive experimental program has been conducted on a LS(1)-0421MOD

airfoil model in the 3'x5' OSU subsonic wind tunnel. Surface pressure distributions were

obtained for 2D baseline and 3D configurations under clean and surface grit conditions.

Several vortex generator configurations were evaluated. The data were taken for steady state

and unsteady conditions. The steady state data included angles of attack from 0/ to 30/ and

Reynolds numbers of 1.0 million. The unsteady conditions were simulated using a face cam

that provided a sinusoidal angle of attack variation with 10/ amplitude for three frequencies

of 0.6 and 1.8 Hz at mean angles of attack of 8/, 14/ and 20/. Surface pressure data were

obtained from six spanwise stations, which were integrated to local coefficients. The

maximum 2D lift coefficient obtained for the 1.0 million Reynolds number was 1.58 at 14.4/

angle of attack. For the 3D case the maximum lift coefficient at the wall was 1.58 at 19.5/

and at the tip was 1.20 at 18.3/. The results showed that the application of the grit roughness

reduces the maximum lift coefficients in all configurations by as much as 50%.

The Flat and Curled vortex generators increased the maximum lift coefficient for both

the 3D tip and wall stations, up to 1.6 and 1.92, respectively. The application of the vortex

generators shifted the stall angle of attack by approximately 30%. A gritted model with the

ii
vortex generators saw an increase in both the maximum lift and stall angle of attack by

approximately 25% in comparison to grit only.

The unsteady maximum lift coefficients were always higher than those for the steady

state up to 60% and showed, generally, large hysteresis loops. The hysteresis loops were

smaller for the 3D wing configuration due to the tip vortex influence, therefore smallest

hysteresis loops occurred at the tip. The Flat and Curled vortex generators removed the

hysteresis loops for all frequencies at 14/ mean angle and significantly reduced the minimum

value of the pitching moment and the pressure drag at stall.

iii
DEDICATION

To Misiek

iv
ACKNOWLEDGMENT

I would like to thank Professor Gerald Gregorek, my adviser, for his guidance and

comments throughout my research. Also Dr. John Lee's help and suggestions through the

many days of testing were constructive and are sincerely appreciated. I would also like to

thank my Graduate Committee for their support. Technical assistance from the staff of the

Aeronautical and Astronautical Laboratory is greatly appreciated.

I would like to thank the National Renewable Energy Laboratories for their continual

support throughout my study at this university.

I also would like to thank Professor Stanley Bruckenstein for his introduction to science

and scientific research. Lastly, my mother's encouragement and willingness to help was a

source of great comfort.

v
VITA

February 25, 1972 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Born - Warsaw, Poland

August 1990 - September 1990 . . . . . . . . . . . . . . . . . . . . . . . Research Support Specialist


State University of New York at Buffalo

August 1992 - September 1992 . . . . . . . . . . . . . . . . . . . . . . . . Research Project Assistant


State University of New York at Buffalo

June 1994 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . B.S., The Ohio State University,


Columbus, Ohio

December 1996 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . M.S., The Ohio State University


Columbus, Ohio

1996-Present . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Graduate Research Assistant


The Ohio State University, Ohio

PUBLICATIONS

1. Janiszewska, J.M., Ramsey, R., Hoffmann M.J., and Gregorek, G.M.: “Effects of Grit
Roughness and Pitch Oscillations on the LS(1)-0417MOD Airfoil”, NREL, Golden,
Colorado, TP-442-7819, January 1996.

2. Ramsey, R., Janiszewska, J.M., and Gregorek, G.M.: Wind Tunnel Testing of Three S809
Aileron Configurations for the Use on Horizontal Axis Wind Turbines”, NREL, Golden,
Colorado, July 1996.

3. Janiszewska, J.M.,Ramsey, R., Hoffmann M.J., and Gregorek, G.M.: “Effects of Grit
Roughness and Pitch Oscillations on the S814 Airfoil”, NREL, Golden, Colorado, TP-442-
8261, May 1996.

vi
4. Ramsey, R., Janiszewska, J.M., and Gregorek, G.M.: “Effects of Grit Roughness and
Pitch Oscillations on the S825 Airfoil”, NREL, Golden, Colorado, October 1998.

5. Janiszewska, J.M., Ramsey, R., Lee, J.D., and Gregorek, G.M.: “Effects of Grit
Roughness and Pitch Oscillations on the L303 Airfoil”, NREL, Golden, Colorado, February
1999.

6. Bruckenstein, S. and Janiszewska, J.M.: "Diffusion Currents to {Ultra}-Micro Electrodes


of Variable Geometries: Ellipsoids, Spheroids and Elliptical Disks", Journal of
Electroanalytical Chemistry, Vol. 538-539, pp 3-12, 2002.

7. Janiszewska, J.M. and Gregorek G.M.: “Three Dimensional Effects on a Wind Turbine
LS(1)-0417MOD Airfoil Model”, NREL, Golden, Colorado, May 2002.

8. Janiszewska, J.M., Gregorek, G.M., and Lee J.D.: “Aerodynamic Characteristics of the
LS(1)-0417 MOD Airfoil Model”, AIAA 2003-0349, 2003 ASME Wind Energy Symposium,
Reno, NV.

9. Janiszewska, J.M., Gregorek, G.M.: “The LS(1)-0417MOD Airfoil Aerodynamic Flow


Characteristics with the Application of Vortex Generators”, AIAA 2004-0661 , 2004 ASME
Wind Energy Symposium, Reno, NV.

FIELDS OF STUDY

Major field: Aeronautical and Astronautical Engineering

vii
TABLE OF CONTENTS
Page

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
DEDICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
ACKNOWLEDGMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
VITA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv
INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
EXPERIMENTAL SETUP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1 Airfoil Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Wind tunnel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Vortex Generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.6 Data Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.7 Test Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
ANALYTICAL METHODS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 3D steady state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
RESULTS AND DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.1 Steady State Data: Clean and Roughness Applied with Vortex Generators . . . . 39
4.1.1 Steady State Integrated Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2 Unsteady Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.2.1 2D Unsteady Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.2.2 2D Unsteady Data with Vortex Generators . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2.3 2D and 3D Unsteady Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.2.4 3D Unsteady Data for All Model Conditions . . . . . . . . . . . . . . . . . . . . . . . 95
4.2.5 Unsteady Integrated Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
LIST OF REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Appendix A: Measured Model Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Appendix B: Summary of Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Appendix C: Permissions for reprinting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

viii
LIST OF FIGURES

Figure Page

1. Steady state and unsteady flows, reproduced with permission from Ref 36. . . . . . . . . 3

2. Unsteady flow characteristics, reproduced with permission from Reference 2. . . . . . . 5

3. LS(1)-0417MOD airfoil model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

4. Model surface irregularities. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

5. LS(1)-0417MOD separated in sections. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

6. PSI modules in the model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

7. LS(1)-0417MOD 3D airfoil model in the 3x5 wind tunnel. . . . . . . . . . . . . . . . . . . . . 13

8. Shaft size comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

9. Roughness pattern . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

10. 3x5 wind tunnel schematic, top view. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

11. 3x5 wind tunnel schematic, side view. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

12. 3x5 wind tunnel oscillation system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

13. Schematic of Type A and B vortex generators. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

14. Schematic of Flat and Curled vortex generators. . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

15. Vortex generators types. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

16. Acquisition setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

17. Theoretical lift and pitching moment comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . 29

ix
18. Theoretical drag comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

19. Gormont at 0.6 Hz, reproduced with permission from reference 44. . . . . . . . . . . . . 32

20. Gormont at 1.8 Hz, reproduced with permission from reference 44 . . . . . . . . . . . . 33

21. Vortex pattern, reproduced with permission from Horner [31]. . . . . . . . . . . . . . . . . 34

22. Tip vortices, reproduced with permission from Horner [31]. . . . . . . . . . . . . . . . . . . 35

23. 3D theory comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

24. Cp distribution for Cl=0.5, clean. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

25. Cp distribution for Cl=1.15, clean. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

26. Cp distribution for Cl=1.0, Flat vortex generators. . . . . . . . . . . . . . . . . . . . . . . . . . . 42

27. Cp distribution for Cl=1.5, Flat vortex generators. . . . . . . . . . . . . . . . . . . . . . . . . . . 43

28. 2D lift coefficient. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

29. 2D drag polar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

30. Lift coefficient at z=0.075. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

31. Pitching moment coefficient at z=0.075 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

32. Lift coefficient at z=0.925 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

33. Pitching moment at z=0.925 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

34. Cp distribution for Cl = 1.0 with grit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

35. Cp distribution for Cl=1.15, grit and VGS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

36. Lift coefficient at z=0.075, grit with VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

37. Pitching moment at z=0.075, grit with VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

38. Lift coefficient at z=0.925, grit with VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

39. Pitching moment at z=0.925, grit with VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

x
40. Integrated lift coefficient, clean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

41. Integrated lift coefficient, Flat VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

42. Integrated lift coefficient, grit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

43. Integrated lift coefficient, grit with VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

44. Total drag polar, clean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

45. Total drag polar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

46. 2D unsteady lift coefficient, clean, f = 1.8H z. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

47. 2D unsteady pitching moment coefficient, clean, f=1.8 Hz. . . . . . . . . . . . . . . . . . . 68

48. 2D unsteady lift coefficient, clean, f = 0.6 Hz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

49. 2D unsteady lift coefficient, LEGR, 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

50. 2D unsteady pitching moment coefficient, LEGR, f = 1.8 Hz. . . . . . . . . . . . . . . . . . 72

51. 2D unsteady lift and pitching moment clean and LEGR, f = 1.8 Hz. . . . . . . . . . . . . 73

52. 2D lift-time comparison, clean and LEGR, f = 1.8Hz. . . . . . . . . . . . . . . . . . . . . . . . 74

53. 2D lift coefficient comparison, clean and LEGR, f = 0.6Hz. . . . . . . . . . . . . . . . . . . 75

54. 2D lift coefficient comparison, clean and LEGR, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . 76

55. 2D lift coefficient comparison, clean and LEGR, f = 0.6 Hz. . . . . . . . . . . . . . . . . . . 77

56. 2D lift coefficient comparison, clean and VGS, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . 78

57. 2D Lift-time, clean and vortex generators, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . . . 80

58. 2D moment coefficient, clean and Flat VGS, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . 81

59. 2D lift coefficient, clean and Flat VGS, f = 0.6 Hz. . . . . . . . . . . . . . . . . . . . . . . . . . 82

60. 2D lift coefficient, clean and Flat VGS, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . . . . . 83

61. 2D lift- time comparison, clean and Flat VGS, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . 84

xi
62. 2D pitching moment coefficient, clean and Flat VGS, f = 1.8 Hz. . . . . . . . . . . . . . . 85

63. Comparison of 2D and 3D lift, clean, f = 0.6 Hz, z=0.075. . . . . . . . . . . . . . . . . . . . 86

64. Comparison of 2D and 3D lift, clean, f = 1.8 Hz, z=0.075. . . . . . . . . . . . . . . . . . . . 87

65. Comparison of 2D and 3D clean moment, wall section, f=1.8 Hz, z=0.075. . . . . . . 88

66. Comparison of 2D and 3D pressure drag, clean, f = 1.8 Hz, z=0.075. . . . . . . . . . . . 89

67. Comparison of 2D and 3D lift, clean, f = 0.6 Hz, z=0.925. . . . . . . . . . . . . . . . . . . . 90

68. Comparison of 2D and 3D pressure drag, clean, f = 1.8 Hz, z = 0.925. . . . . . . . . . . 91

69. Comparison of 2D and 3D lift, clean, f = 1.8 Hz, z = 0.925. . . . . . . . . . . . . . . . . . . 92

70. Comparison of 2D and 3D pitching moment, clean, f = 1.8 Hz, z = 0.925. . . . . . . . 93

71. Comparison of 2D and 3D clean pressure drag, f = 1.8 Hz, z = 0.925. . . . . . . . . . . 94

72. 3D steady and unsteady comparison, f=1.8 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . 95

73. 3D steady and unsteady comparison, f = 1.8 Hz, z = 0.925. . . . . . . . . . . . . . . . . . . . 96

74. 3D lift clean and VGs, f = 0.6 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

75. A time comparison of 3D clean and VGs, f=0.6 Hz, z = 0.075. . . . . . . . . . . . . . . . . 98

76. 3D moment clean and VGs, f = 0.6 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . . . . . . . 99

77. 3D pressure drag clean and VGS, f = 0.6 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . 100

78. 3D lift clean and VGs, f = 1.8 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

79. 3D moment clean and VGs, f = 1.8 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . . . . . . 102

80. 3D lift clean and VGs, f = 1.8 Hz, z = 0.925. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

81. 3D moment clean and VGs, f = 1.8 Hz, z = 0.925. . . . . . . . . . . . . . . . . . . . . . . . . . 104

82. 3D k/c=0.0019 Cl vs ", ±5.5/, Tred~0.022 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

83. 3D lift clean and Curled VGs, f = 1.8 Hz, z = 0.925. . . . . . . . . . . . . . . . . . . . . . . . 106

xii
84. 3D lift grit and VGs with grit, f = 1.8 Hz, z = 0.075. . . . . . . . . . . . . . . . . . . . . . . . 107

85. 3D lift grit and grit with VGS, f = 0.6 Hz, z = 0.925. . . . . . . . . . . . . . . . . . . . . . . . 108

86. 3D induced angle of attack, clean and VGs, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . 109

87. 3D integrated lift, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

88. 3D integrated drag, f = 1.8 Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

xiii
LIST OF TABLES

Table Page

A1. LS(1)-0417MOD 3D Surface Pressure Taps, Non-Dimensional Coordinates for

z=0.258, z=0.592, and z=0.925 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

A2. LS(1)-0417MOD 3D Surface Pressure Taps, Non-Dimensional Coordinates for

z=0.075, z=0.408, and z=0.742 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

B1. LS(1)-0417MOD 3D Steady State Parameters Summary for 2D . . . . . . . . . . . . . . 127

B2. LS(1)-0417MOD 3D, 2D Unsteady, Clean, ±5.5/ . . . . . . . . . . . . . . . . . . . . . . . . . 127

B3. LS(1)-0417MOD 3D, 2D Unsteady, LEGR, ±5.5/ . . . . . . . . . . . . . . . . . . . . . . . . 128

B4. LS(1)-0417MOD 3D, 2D Unsteady, Clean, ±10/ . . . . . . . . . . . . . . . . . . . . . . . . . 128

B5. LS(1)-0417MOD 3D, 2D Unsteady, LEGR, ±10/ . . . . . . . . . . . . . . . . . . . . . . . . . 129

B6. LS(1)-0417MOD 3D Steady State Parameters Summary for 3D . . . . . . . . . . . . . . 130

B7. LS(1)-0417MOD 3D Steady State Parameters Summary for 3D with VGs . . . . . 131

B8. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.075 . . . . . . . . . . . . . . . . . . . . . 131

B9. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.075 . . . . . . . . . . . . . . . . . . . . 132

B10. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.258 . . . . . . . . . . . . . . . . . . . . 132

B11. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.258 . . . . . . . . . . . . . . . . . . . 133

B12. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.408 . . . . . . . . . . . . . . . . . . . . 133

B13. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.408 . . . . . . . . . . . . . . . . . . . 134

xiv
B14. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.592 . . . . . . . . . . . . . . . . . . . . 134

B15. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.592 . . . . . . . . . . . . . . . . . . . 135

B16. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.742 . . . . . . . . . . . . . . . . . . . . 135

B17. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.742 . . . . . . . . . . . . . . . . . . . 136

B18. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.925 . . . . . . . . . . . . . . . . . . . . 136

B19. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.925 . . . . . . . . . . . . . . . . . . . 137

B20. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.075 . . . . . . . . . . . . . . . . . . . 137

B21. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.075 . . . . . . . . . . . . . . . . . . . 138

B22. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.258 . . . . . . . . . . . . . . . . . . . 138

B23. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.258 . . . . . . . . . . . . . . . . . . . 139

B24. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.408 . . . . . . . . . . . . . . . . . . . 139

B25. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.408 . . . . . . . . . . . . . . . . . . . 140

B26. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.592 . . . . . . . . . . . . . . . . . . . 140

B27. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.592 . . . . . . . . . . . . . . . . . . . 141

B28. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.742 . . . . . . . . . . . . . . . . . . . 141

B29. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.742 . . . . . . . . . . . . . . . . . . . 142

B30. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.925 . . . . . . . . . . . . . . . . . . . . 142

B31. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.925 . . . . . . . . . . . . . . . . . . . 143

xv
LIST OF SYMBOLS

A/C Alternating current


AR Aspect ratio
b Span
c Chord of the airfoil
CD Total drag coefficient
Cdi Induced drag coefficient
Cd min Minimum drag coefficient
Cdp Pressure drag coefficient
Cdw Wake drag coefficient
Cl Local lift coefficient
ClREF Reference lift coefficient
ClUNST Unsteady predicted lift coefficient
CL Total lift coefficient
Cl max Maximum lift coefficient
Cl dec Maximum lift coefficient at a decreasing angle of attack
Cm, Cm ¼ Pitching moment coefficient about the quarter chord
Cm dec Pitching moment coefficient at maximum lift with the angle of attack decreasing
Cm inc Pitching moment coefficient at maximum lift with the angle of attack increasing
Cmo Pitching moment coefficient at zero lift
Cp Pressure coefficient
D Drag
e Efficiency factor
f Frequency
h Wind tunnel test section height
Hp Horsepower
Hz Hertz
K1 Unsteady aero empirical factor for αD < 0
k/c Grit particle size divided by airfoil model chord length
L Lift
p Pressure
psw Static pressure in the wake
ptw Total pressure in the wake
p4 Freestream static pressure
q Dynamic pressure
qws Dynamic pressure in the wake

xvi
q4 Freestream dynamic pressure
Re Reynolds number
t Time
U4 Corrected free stream velocity
V Velocity
w Downwash distribution across the span
x Coordinate parallel to model reference line
y Coordinate perpendicular to the reference line
z Span location
" Angle of attack
"Cl=0 Angle of attack at 0 lift coefficient
" dec Decreasing angle of attack
"i Induced angle of attack
" inc Increasing angle of attack
"REF Reference lift coefficient
α Time derivative of angle of attack
) Incremental angle of attack, lift or drag
(1 Stall delay function at low pitch rates
(2 Stall delay function at high pitch rates
' Vorticity distribution across the span
'0 Initial vorticity
, Tunnel solid wall correction scalar
,sb Solid blockage correction scalar
,wb Wake blockage correction scalar
7 Body-shape factor (0.305 used)
D Density
F Tunnel solid wall correction parameter
T red Reduced frequency, B f c/U4
< Kinematic viscosity

xvii
CHAPTER 1

INTRODUCTION

Time-dependent three-dimensional flow fields continue to be the most challenging

aspects in aerodynamics with very few solution techniques (in CFD) and even less

experiments to help model the physics. The current study is unique in the detailed

documentation of the behavior of an oscillating finite wing (or turbine blade) together with

the related steady-state in both two-dimensional and three-dimensional model configurations.

The pressure distributions across the span provide information on the effect of the tip vortex

in the different types of flows. In addition, the influence of vortex generators as a passive

boundary-layer control devices was investigated.

Airfoil flow fields are currently fairly well documented (experimentally), understood and

predictable for steady-state two-dimensional cases and even with the advent of deep-stall and

supercritical flows the basic physics of such problems have been characterized by experiment

and, to some extent, by theory [47]. The application of the two dimensional airfoil

characteristics to the finite wing, propeller and turbine has been successful for some time.

The introduction of unsteadiness (by oscillation or abrupt attitude change) changes the

difficulties in analysis and experiment by at least two orders of magnitude for the airfoil

(two-dimensional) and much more for the wing (three-dimensional).

1
Oscillating Airfoil

For some time the unsteady problem has been understood as characterized by the two

aspects of the potential flow and the boundary layer. In fact, for the majority of the problems

in applied aerodynamics the potential flow can be described as quasi steady, that is, the time-

dependent disturbance is slow enough that the potential flow would be "instantaneously

steady-state". However, the response of the boundary layer is much slower and important

features of the boundary layer such as transition and separation become altered (i.e. delayed).

The influence of the boundary layer on the potential flow (displacement, stall) are well-

known for steady flows and when these features enter the unsteady problem the result is that

the potential flow (beyond the layer) departs from quasi steady one as the boundary layer

dominates the field. Thus, a rapidly pitching airfoil or wing experiences a considerable

increase in lift beyond its steady-state value as the separation of the boundary layer is

delayed; similarly the return to normal attached flow is delayed. Thus the observed hysteresis

loop is generated; although poorly understood and generally unpredictable, it has been well-

documented for many cases experimentally and the results applied empirically for the

progress of aerodynamics.

For example, in 2D tests the normal force increased 20% and the chordwise force

increased 40% on a NACA 0018, tested under pitching conditions [1]. Also as can be seen

in figure 1, reproduced from Reference 36, the unsteady lift coefficient is approximately 25%

higher than the steady state value. The unsteady quarter chord pitching moment and drag

coefficients are usually not affected as much by the oscillating behavior of the airfoil. The

moment especially is close to the steady state values, except a sharp decrease occurs near the

2
stall point [36]. The drag coefficient increases with the unsteady motion in comparison to

the 2D steady state [36]. The large increase in the normal or axial forces will cause an

additional stress on the wing or wind turbine blade. Also the boundary layer in an unsteady

case is affected by the type and speed of the oscillation. In the upswing the boundary layer

usually stays attached longer than in the steady state case and therefore higher lift occurs.

However, after stall in the downswing the boundary layer does not reattach as quickly, and

a thicker boundary layer exists for the same angles of attack as on the upswing.

Figure 1. Steady state and unsteady flows, reproduced with permission from Ref 36.

Even in the quasi-steady flow the aerodynamic characteristics differ greatly between the

upswing and downswing of the airfoil. These differences can cause high blade stress, blade

vibrations and even control loss [26]. The hysteresis loops in lift, pitching moment and drag

3
are also a result of the inability of the boundary layer to follow the same path on the upswing

and downswing thus affecting the potential flow. For example, the lift coefficient can vary

anywhere between 10 to 60% between the upswing and downswing. The reattachment point

is not easily predicted, since it varies with the Reynolds number, grit roughness and

application of vortex generators even for the same model shape. The hysteresis size and

shape depend on the airfoil shape, the frequency of the oscillation, the mean angle and the

Reynolds number.

In the upswing the stall on the airfoil is delayed and as a result a higher maximum lift

coefficient is obtained. It has been suggested that the delay in the stall is due to a later

transition from the laminar to the turbulent boundary layers. The laminar boundary layer

dominates the flow characteristics for small angles of attack on the upswing [28]. With an

increase in angle of attack close to stall the boundary layer separates from the airfoil surface

which leads to the formation of the dynamic stall vortex (DSV). In addition to the DSV,

another clockwise rotating vortex, the trailing edge vortex (TEV) is formed as the airfoil

reaches high angles of attack. Together the vortex structures produce large turbulence which

is moved to the wake as the angle of attack decreases [2]. These flow characteristics are

shown in figures 1 and 2, reproduced from References 36 and 2, respectively. In figure 2 the

actual streamlines are shown and they represent a thinning in the boundary layer on the

upswing and an increase in thickness on the downswing. The flow field changes slightly as

the frequency of the oscillation in increased. For higher frequencies the maximum lift

coefficient increased more than for the lower frequencies, most likely due to the slow

turbulence propagation in the outer part of the boundary layer [28]. In an experiment

4
performed in France, the boundary layer thickness was shown to decrease as the reduced

frequency increased [28].

Figure 2. Unsteady flow characteristics, reproduced with permission from Reference 2.

The unsteady flow affects the effectiveness of the high-lift devices, such as vortex

generators or flaps. The effects of the dynamic stall on vortex generators are not well

5
understood and little data can be found that deals with unsteady flow conditions. The flow

characteristics vary highly depending on the shape of the device and the type of oscillation.

The next step in introducing three-dimensionality (i.e. the finite wing, rotor, turbine

blade) elevates the complexity of the problem even further. The influence of the downwash

and the characteristics of the tip vortex system are now forced to become unsteady by the

wing boundary layer so that the interactions become more complex.

This study was aimed at investigating a simple three dimensional case: that of a

untwisted wing with an aspect ratio of 3.3 and a constant chord (spanwise) in a sinusoidal

oscillation. However, the oscillation was designed to force the wing into deep stall as well

as unstalled. Data were taken over the entire span and the separate influences of surface

roughness and of vortex generators were documented. In order to provide as much

understanding as possible, a complete base of steady flow and of two-dimensional flow was

generated and, wherever possible, theoretical predictions were considered. While the 2D

unsteady flow is difficult to predict and understand since it depends on the type of oscillation,

the frequency or even the model surface finish, the 3D flow includes effects from the tip

vortex and the twist and taper of the platform, which have to be taken into account. In the

3D steady state the maximum lift coefficient is usually lower than in the 2D case but the stall

angle of attack increases, due to the change in the lift curve slope [31]. The lift curve slope

is lower closer to the tip and increases further away from the tip; in theory it would become

the 2D value at infinity. According to the data from a delta wing test under unsteady

conditions, the vortex breakdown is large and controls the high maximum lift and the

hysteresis behavior [29]. Additional flow characteristics have to be examined when dealing

6
with 3D unsteady flow, including the effect of rotation, which causes a delay in stall and

therefore a higher lift than in 2D [39]. Additionally even in quasi steady flow Björck found

that for 3D the behavior of the wing is not easily predicted and the measured data show a

“unsteadiness” inherent to the type of oscillation.

Present Work

The main purpose of this study was to investigate a three dimensional wing or wind

turbine configuration with and without vortex generators for both steady state and unsteady

flows. Data for this type of problem are rarely available in the literature. The numerous

pressure distributions obtained across the span provide an insight into the flow characteristics

around a three dimensional airfoil model. The data available in the literature generally do

not contain pressure distributions, while those that are included are very limited and usually

do not include sufficient spanwise data making the validation of the theoretical codes

difficult. These current tests provide pressure data for both baseline two dimensional and

three dimensional model configurations, i.e. clean model condition without roughness or

vortex generators. The airfoil model for this study was designed as a simplified wing, with

no twist and a constant chord to isolate the 3D effects from the more complicated flow fields

of a general wing. The subboundary layer vortex generators are popular in the literature,

especially in controlling the boundary layer on the test section walls [43]. In this study a

multitude of different vortex generators shapes were tried in order to increase the maximum

lift coefficient with a minimum drag penalty and to control the hysteresis loop in unsteady

flows. Another part of the research dealt with the surface degradation of the wind turbine

7
blades and wings due to insect accumulation or age of the surface, this data are mostly only

available for small grit strips (like zigzag tape trip strips) but not for a full severe

contamination [37]. Both types of data, with slight and severe roughness are need to

understand the effect of grit on the aerodynamic characteristics especially after the vortex

generators are applied to the surface of the wings. The surface roughness was simulated for

these experiments was reproduced from an actual pattern which occurred on a wind turbine

in the field. The main focus of this study is unsteady flow characteristics of the 3D wing

section with both the vortex generators and surface roughness, since little experimental data

was found in the literature. The data presented in the literature are mostly limited to 2D

configuration and also in type of oscillation, the amplitude, the frequency or the mean angle

of attack.

All the experiments performed for this study confirm that it is difficult to predict the

flow response to the vortex generators, especially in 3D unsteady flow. A slight change in

the shape of the actual vortex generator or its position and the flow response changes

dramatically, even if the previous data suggest that the response should be better. The only

vortex generators presented in this text are ones that gave very consistent results and were

not affected as much by changes in position or shape.

8
CHAPTER 2

EXPERIMENTAL SETUP

2.1 Airfoil Model

The LS(1)-0417MOD airfoil, which is 17% thick, was used for these experiments. It is

an unclassified NASA profile, designed for general aviation with an original design lift

coefficient of 0.4. The trailing edge on this airfoil has been modified to a finite thickness for

easier manufacturing. Figure 3 shows the airfoil shape.

Figure 3. LS(1)-0417MOD airfoil model.

A 457 mm (18 in) constant chord model was designed and manufactured using the

LS(1)-0417MOD profile. The airfoil model was made in four interchangeable sections. The

model was constructed from a nine layer composite lay up of alternating fiberglass and

carbon fiber over ribs. The main load bearing member was a 38 mm (1.5 in) diameter steel

9
tube 213 mm (84 in) long which passed through the model quarter chord station. The tube

extended through the whole height of the tunnel for both wind tunnel for both the 2D and 3D

tests, since it has little effect on the flow around the airfoil model but adds structural stability

for the unsteady test. Ribs and end plates were used to transfer loads from the composite

skin to the steel tube. The final surface was filled, painted and wet sanded to attain given

coordinates within a requested tolerance of ±0.25 mm (±0.01 inches). The completed model

was measured at the three spanwise locations using a Sheffield-Cordax coordinate

measurement machine. Figure 4 shows the normal differences between the measured and

desired coordinates. The tolerance limit for accepting the manufactured model was ±1mm

difference, therefore the model was found to be within the tolerance.

Upper surface
Lower surface

1
Ydesired-Ymeasured (mm)

-1

-2
0 115 230 345 460

X position (mm)

Figure 4. Model surface irregularities.

10
Figure 5. LS(1)-0417MOD separated in sections.

The model with four sections separated is shown in figure 5. The alignment of the four

pieces was guaranteed by pins running through the length of the model and additional set

screws from the sides that held the individual parts of the model to the support shaft. Only

one section was instrumented with two rows of pressure taps (the section with the hatch).

The coordinates of the pressure taps are listed in Appendix A. The 76 surface pressures

were measured simultaneously by the PSI system and then reduced to the individual pressure

distributions. The three sections were interchangeable, making it possible to measure

pressures on the model in six different locations across the span in both the 2D and 3D

configurations. The left hand section was manufactured to obtain 2D baseline data. The 3D

model had a span of 762 mm (30 in). As the sections were moved the data were obtained at

span setting of 0.075, 0.258, 0.408, 0.592, 0.742 an 0.925; based on a 3D span of 762 mm

(30 in). A small endcap was used with the model in the 3D case to remove the sharp edges

11
from the tip. The end cap was 16 mm (0.65 in) height, and that length was not added to the

model span measurement for simplicity.

Figure 6. PSI modules in the model.

To minimize pressure response times, which is important for the unsteady testing, the

length of the surface pressure tap lead-out lines had to be as short as possible. Consequently,

a compartment was build into the model and the pressure scanning modules were installed

inside. This compartment could be accessed through a panel door fitted flush with the model

contour on the lower surface. Figure 6 shows the pressure scanning modules (PSI modules)

installed inside the model. Three modules were used for the test program with each

measuring up to 32 pressures. Two modules had a range of ±2.5 PSID and other one ±5

PSID. The surface pressure lead-out lines were hooked up in no particular order to the PSI

12
modules and a tap file contained the hook up information and when used with the reduction

program arranged the measured pressures in correct order.

Figure 7. LS(1)-0417MOD 3D airfoil model in the 3x5 wind tunnel.

Figure 7 shows the model including the endcap installed in the 3x5 wind tunnel prepared

for the 3D testing with the pressure tapped section on the bottom. As can be seen from the

figure, the steel shaft extends through the top of the tunnel for structural stability.

Early tests were conduced to see if the shaft in that position would affect the flow

characteristics. An outside tube was placed on the shaft, so that the resulting shaft was twice

the diameter of the original (i.e. approximately 3" in diameter, original is 1.5"). The lift

13
curves at the tip station from some of those tests are presented in figure 8. The effect of the

pipe was considered to be minimal, especially for the smaller shaft case. There was no

change in the maximum lift coefficient and the stall angle of attack was approximately the

same. There was a slight change in the shape of the lift curve slope and the stall

characteristics, due to the interference produced by the pipe. For both cases stall occurred

at approximately the same angle of attack and the lift after that point was exactly the same.

Thus it was concluded that the pipe, needed for structural integrity, produced only negligible

interference on the wing.

1.5

1.0
Lift Coefficient

Original diameter
2 times the diameter
0.5

z=0.925
Steady State
Clean
Re=1.0 million

0.0
0 10 20 30

Angle of Attack, deg

Figure 8. Shaft size comparison.

14
Figure 9. Roughness pattern

For test cases involving roughness, a standard roughness pattern developed for the

National Renewable Energy Laboratory airfoil program was used. It was generated using a

molded insect pattern taken from a wind turbine in the field. The particle density was

5 particles per cm2 (35 particles per square inch) in the middle of the pattern, thinning to

1.25 particles per cm2 (8 particles per square inch) at the edge of the pattern. Figure 9 shows

the roughness pattern. The pattern was repeatedly cut into a steel sheet 102 mm (4 in) wide

and 91 cm (3 ft) long with holes just large enough for one grain of grit. Based on average

particle size from the field specimen, standard #40 lapidary grit was chosen for the roughness

elements, giving a particle height to chord ratio of 0.009 for a 457 mm (18 in) chord model.

15
The use of the template and double sided tape created a repeatable leading edge grit

roughness (LEGR) pattern every time.

2.2 Wind tunnel

Figure 10. 3x5 wind tunnel schematic, top view.

Tests were conducted in the Ohio State University Aeronautical and Astronautical

Research Laboratory 3x5 Subsonic Wind Tunnel. Figures 10 and 11 show the schematics

of the wind tunnel, the top and the side views, respectively.

16
Figure 11. 3x5 wind tunnel schematic, side view.

This open circuit wind tunnel has a velocity range of 0-55 m/s (180ft/s) produced by a

2.4 m (8 ft) diameter, six-bladed fan. The fan is belt driven by a 93.2 kw (125 hp) three face

a.c. motor connected to a variable frequency motor controller. Nominal test section

dimensions are 1 m (39 in) high by 1.4 m (55 in) wide by 2.4 m (96 in) long. The 457 mm

(18 in) chord airfoil model was mounted vertically in the test section. A steel tube through

the quarter chord of the model was used to attach the model to the tunnel during testing. An

angle of attack potentiometer was fastened to the model at the top of the tunnel as shown in

figure 11. The steady state angle of attack was adjusted with a worm gear drive and unsteady

angle of attack was controlled by the shaker; both were attached to the model strut below the

tunnel floor.

17
Figure 12. 3x5 wind tunnel oscillation system.

2.3 Oscillation System

For the pitch oscillating tests a shaker system was used, seen in figure 12. It incorporated

a face cam and a follower arm attached to the model support shaft below the wind tunnel

floor. The choice of the cam governs the type and amplitude of the wave form produced.

Sine wave forms having amplitudes of ±5.5/ and ±10/ were used. The wave form was

defined by the equation

(2.3.1)

where A is the respective amplitude. The shaker system was powered by a 5 hp a.c. motor

with a variable line frequency controller. The oscillating frequency range was 0.1 - 2.0 Hz.

Three different frequencies were used for these test: 0.6, 1.3 and 1.8 Hz.

18
2.4 Vortex Generators

Figure 13. Schematic of Type A and B vortex generators.

Many different shapes and sizes of vortex generators were designed and tested for this

exploratory study in an attempt to find the correct ones for increasing lift and decreasing

drag. Based on the data from Reference 43, a set of sub-boundary layer vortex generators

was made, but for this model they showed very little change in lift and almost no change in

the other aerodynamic coefficients, contrary to the results of that paper. According to

Reference 43 the sub-boundary layer vortex generators are ideal, since they increase lift

without the drag penalty.

19
The four types described here were the best from each phase of the program. The vortex

generators were positioned at x/c from 0.07 to 0.5 in attempt to determine the optimum

configurations and locations with regard to the maximum lift and minimum drag.

Type A vortex generators were used as a baseline and Type B was derived from them as

shown in figure 13. Type A were a molded set of plastic vortex generators and were

approximately 0.2" high with a high base which made them actually taller, more like 0.25"

and were 0.38" apart. Type B was made from thin brass sheets, like all the others later, and

had a height of 0.15". Additionally this type was spaced slightly closer together at 0.25"

apart. The angle of the vortex generator to the flow was kept fixed for all the types, since it

was found to be the best through a series of tests. Both Types A and B were used to

determine the optimum position of the vortex generators. It was found to be from 0.07 chord

at the tip and 0.15 chord at the root (wall). This slanted arrangement was used for all 3D

testing. For 2D testing the vortex generators were positioned at a average point, 0.11 chord

to obtain a average lift and wake drag for the configurations.

20
Figure 14. Schematic of Flat and Curled vortex generators.

The Flat and Curled vortex generators were developed later in the study and are shown

in figure 14. The Flat vortex generators are a replica of Type A other than the platform shape

has been changed to almost rectangular. From looking at the earlier data and old reports, it

seemed like the rectangular area vortex generators could work more efficiently. The Curled

vortex generators developed out of the idea that the slight curl will help the vortex on the tip

of the vortex generators interact with the model cross flow. The photograph in figure 15

shows Type A and B on the top and Flat and Curled on the bottom.

21
Figure 15. Vortex generators types.

2.5 Data Acquisition

Data were acquired and processed from 76 surface pressure taps, four individual tunnel

pressure transducers, an angle of attack potentiometer, a wake probe and tunnel

thermocouple. The data acquisition system included an Pressure System Incorporated (PSI)

data scanning system controlled by a PC. The PSI system included a 780B Data Acquisition

and Control Unit (DACU), 780B Pressure Calibration Unit (PCU), 81-IFC scanning module

interface, two 2.5 psid and one 5 psid pressure scanning modules (ESPs), one 20 in water

column range pressure scanning module and a 30 channel Remotely Addressed Millivolt

Module (RAMM-30). Figure 16 shows the acquisition setup.

22
Figure 16. Acquisition setup.

Four individual pressure transducers read: tunnel total pressure, tunnel north static

pressure, tunnel south static pressure, and wake dynamic pressure. These transducers were

bench calibrated using a water manometer to determine their sensitivities and offsets. Those

values were entered into the data acquisition and reduction program so the transducers could

be shunt resistor calibrated before each series of wind tunnel runs.

The rotary angle of attack potentiometer of 0.5% linearity was calibrated during the

tunnel pressure transducers shunt calibration. The angle of attack calibration was

accomplished by taking voltage readings at known angles. This calibration method gave

angle of attack readings within ±0.25/ over the entire angle range. The wake probe position

potentiometer is a linear potentiometer and it was also calibrated during the shunt calibration

of the tunnel pressure transducers.

23
Calibration of the four ESPs was done simultaneously by applying known regulated

pressures to the ESPs and reading the output voltages from each pressure sensor. From these

values, the DACU calculated the calibration coefficients, and stored them internally until the

coefficients were requested by the controlling computer. Frequent on-line calibrations were

performed during a run set since the outputs of the ESPs tended to drift with temperature

changes.

For steady state cases, the model was set to an angle of attack and the tunnel conditions

were adjusted to the desired Reynolds number. Pressure measurements from the airfoil

surface taps and the flow conditions were acquired. The angles of attack were always set in

the same progression - from 0/ to -20/ then from 0/ to +40/. For 2D steady state the wake

data were acquired for angles of attack where the wake was relatively narrow.

For model oscillating cases, the tunnel conditions were set while the model was

stationary at the particular mean angle of attack. The model then was allowed to oscillate

through few cycles to establish the flow field, and finally the model surface pressure and

tunnel condition data were taken. For each run condition, 80 data scans were acquired, that

resulted in approximately 1/ steps in the data set.

2.6 Data Reduction

The ambient pressure was manually entered into the computer and updated regularly. Its

value, as well as the measurements from the tunnel pressure transducers and the tunnel

thermocouple, were used to calculate tunnel airspeed. As a continuous check of readings,

24
the tunnel total and static pressures were read by both the tunnel pressure transducers and the

20-inch water column ESP.

A typical steady state data point was derived by acquiring 10 data scans of all channels

over a 10 second window at each angle of attack and tunnel condition. The reduction portion

of the program processed each scan to coefficient (Cp, Cl, Cm¼, and Cdp) using the measured

surface pressure voltages, calibration coefficients, tap locations and wind tunnel conditions,

and then averaged them into one set. A data set was then corrected for the effects of solid

tunnel walls.

Corrections due to solid tunnel sidewalls as described by Pope and Harper were applied

to the wind tunnel data for both 2D and 3D [45]. No other 3D corrections were applied to

the data. Tunnel conditions are represented by the following equations, where the subscript

u denotes uncorrected values:

(2.6.1)

(2.6.2)

(2.6.3)

Airfoil aerodynamic characteristics are corrected by:

(2.6.4)

(2.6.5)

(2.6.6)

(2.6.7)

where

(2.6.8)

25
(2.6.9)

(2.6.10)

(2.6.11)

In order to determine the drag accurately, the wake momentum method was used. Model

wake data were taken using a horizontally traversing pitot-static probe. The measured

pressures were used to calculate drag coefficient using a form of the Jones equation derived

from Schlichting [46].

(2.6.12)

This equation assumes static pressure at the measurement site is the free-stream value. The

integration was done automatically using chosen end points. This technique was only

possible for steady state, unstalled conditions and beyond them the less accurate pressure

drag was calculated from the data acquired.

For pitch oscillation cases, model surface pressures were reduced to lift, moment and

pressure drag coefficients. There was no calibration available for unsteady model pitch

conditions, so the unsteady pressure data were not corrected for effects due to time dependent

pitching or solid tunnel walls. Also, the wind tunnel contraction pressures (used for steady

state cases) could not be used to calculate instantaneous free stream conditions due to slow

response times. The tunnel conditions were obtained from a total pressure probe, and the

average of opposing static taps at the test section entrance.

26
2.7 Test Matrix

The test was designed to provide two-dimensional(2D) and three-dimensional(3D) data

for steady state and pitch oscillating model of a wind turbine blade with and without vortex

generators. Steady state data, required for comparisons, were acquired at Reynolds numbers

of 0.75, 1, and 1.25 million with and without LEGR. The angle of attack increment was two

degrees when -20/ < " < +10/ or +20/ < " <+40/ and one degree when +10/ < " < +20/ .

Wake surveys were conducted to find total airfoil drag over an angle of attack range of -10/

to +10/ for 2D. Unsteady data were taken for Reynolds numbers of 0.75, 1, and 1.25 million.

Sine wave cams having amplitudes ±5.5/ and ±10/ were used for pitch oscillations with the

mean angles of 8/, 14/, and 20/ and for frequencies of 0.6, 1.3, and 1.8 Hz. A Reynolds

number of 1 million was selected for the vortex generator tests. The different types of vortex

generators were applied to the 2D and 3D model configurations for both steady state and

pitch oscillating cases. The angle of attack ranged from 0/ to 30/ for the vortex generators

cases.

27
CHAPTER 3

ANALYTICAL METHODS

3.1 2D steady state

The lift, moment and drag coefficients for a 2D steady state airfoil model can be

accurately predicted using any of the available codes, such as NC State, MACARFA and

Eppler providing there is no unusual boundary layer behavior, such as stall, bubbles and that

the entire flow field is sub critical (local much number is less than 1). These panel method

codes divide the airfoil surface into sections and apply a inviscid linear equations to each

piece. The flow is approximated by a uniform flow and a vortex centered on a panel, so that

each panel is a sheet of vorticity. The correct strength of the individual vortices is found

through the boundary conditions. The two boundary conditions are: there is no flow through

the surface of the airfoil and the Kutta condition, which states that there is no flow around

the trailing edge.

The obtained 2D wind tunnel data were compared with results from the NC State code

[48]. In addition to the panel method, the code calculates the boundary layer characteristics

on the airfoil to obtain the drag. At least two iterations are made, with the boundary layer

displacement thickness having been added to the airfoil surface. As shown in figure 17, the

28
lift coefficient is well predicted for the low angles of attack from 0/ to 10/. The stall region

is not predicted well, since the code does not account for separation. The prediction of the

pitching moment coefficient about the quarter chord shown together with the lift is good for

all the angles of attack. The predicted drag coefficient, shown in figure 18 is excellent, the

discrepancies occur closer to stall, where the code underestimates drag due to large flow

separation.

e o
2.1 0.5

0.4
1.8
0.3

1.5 0.2

Pitching Moment Coefficient


0.1
Lift Coefficient

1.2
-0.0

0.9
-0.1

-0.2
0.6
2D data - C
l
NCS - C -0.3
l
2D data - C
0.3 NCS - C
m

m
-0.4
Steady State
0.0 -0.5
0 10 20 30

Angle of Attack, deg

Figure 17. Theoretical lift and pitching moment comparison.

29
0.0 5

0.0 4
Drag Coefficient

0.0 3

0.0 2

2D data
N C S Theory
0.0 1
S teady S tate

0.0 0
0 2 4 6 8 10 12 14 16 18 20

A ngle of A ttack , deg

Figure 18. Theoretical drag comparison.

This study did not center on theoretical code predictions but on the actual experimental

aerodynamic coefficients and the effect vortex generators have on unsteady flow. The CFD

programs were not used to obtain data in the separated region or for the model with the

vortex generators. The predictions for the vortex generators would be difficult since even

in experiments a tiny change can make a huge difference in the results. For a CFD solution

both the separated flow after stall and behind the vortex generators would have to be

modeled and the Navier-Stokes equation would have to be solved. The main problem with

modeling the vortex generators effects is that they are difficult to predict on the actual airfoil

model.

30
3.2 2D Unsteady

For unsteady 2D flows, there are a few time dependent prediction methods, which

include unsteady CFD or the Gormont approximate method. The CFD requires time in

setting up, for example in generating grids, turbulence models, may provide pressure

distribution in time. The turbulence generated by the oscillating airfoil has to be modeled

in some fashion [9,12,17]

The Gormont method is an empirical method for predicting the lift coefficient with

respect to the angle of attack by using 2D steady state data [50]. The method was developed

by finding a correlation between the unsteady hysteresis loop and the steady state. Gormont

developed this approach in order to apply it to lift, but it can be extended to both moment and

pressure drag. The method is very simple and follows this formulation

α REF = α − K1∆ Dynamic stall


(3.2.1)

Where )Dynamic stall depends on the angle of attack where the stall occurs, before stall it is

defined,

cα
∆ Dynamicstall = γ 1 sign(α ) (3.2.2)
2V

After stall this formula is used,

 cα  cα cα   α


∆ Dynamicstall =  γ 1 + γ 2  −  
  (3.2.3)
 2V break  2V 2V break   α

31
The lift coefficient is predicted by using the following equation,

 
( )
ClREF
ClUNST =   α − α C = 0 (3.2.4)
 α REF − α Cl = 0  l

The maximum thickness to chord ratio is used to determine the break point for stall and (1

and (2 are calculated from test Mach number.

Figure 19. Gormont at 0.6 Hz, reproduced with permission from reference 44.

32
Figure 20. Gormont at 1.8 Hz, reproduced with permission from reference 44

As can be seen from figure 19 taken from reference 44, the predicted data agreed quite

well for the low frequency case of 0.6 Hz for this airfoil. In this case the oscillations are

slow and the boundary layer on the airfoil has time to adjust. The hysteresis loops for the

low frequency case are usually small and the data are close to the 2D steady state, so that

predictions can be close to the actual measured values. Once the frequency of the

oscillations is increased, the prediction becomes much worse, as seen from figure 20.

Normally the Gormont method predicts slightly lower maximum lift coefficient and stall

usually occurs sooner than in experiments. Also the hysteresis loops that are predicted have

a slightly different shape. They exhibit a sharper stall a few degrees after the maximum lift

33
coefficient is reached. The lift in the predictions increases much faster than in the

experiment and reattaches to the upswing sooner by approximately 3/. It is possible to

extend the Gormont method to the drag and moment coefficients for the unsteady cases,

since both exhibit hysteresis loops.

The Gormont method could be used for cases with vortex generators, even though the

correlation between steady state and hysteresis loops is difficult to find and also that

correlation would contain some error due to the unpredictable response to the vortex

generators.

Figure 21. Vortex pattern, reproduced with permission from Horner [31].

34
Figure 22. Tip vortices, reproduced with permission from Horner [31].

3.3 3D steady state

Flow over a finite wing is much more complicated than that of a 2D wing, since a tip

vortex now exists. The theory of the 3D finite wing was developed by Lanchester and

Prandtl in 1911. According to their theory there is a discontinuity surface which exists and

extends past the trailing edge with that surface being dependent upon the aspect ratio [49].

The flow characteristics are a summation of the uniform flow, the vortex flow over the wing

and the trailing edge vortex. A tip vortex flow is generated when the flow from the lower

surface is coming over the tip of the wing to the upper surface and it is driven by the pressure

difference between the wing surfaces, as seen in figure 21 and 22. The downwash is a new

feature in the flow, and is defined as a downwards deflection of the flow due to the tip

35
vortices. On a 3D wing the lift, moment and drag coefficients vary across the span and are

now a function of position, because the downwash produces an induced angle of attack.

From Prandtl’s theory [48], the lift of the wing can be found from this simple integral,

B/2

L = ρV ∫ Γ ( y)dy
− B/2
(3.3.1)

Where '(y) is the local (spanwise) vorticity distribution on the wing. The drag on the wing

is a summation of the viscous and induced drag. The viscous drag can be found from a

boundary layer analysis and the induced drag is defined as follows,

B/2

D= ρ ∫ w( y)Γ ( y)dy
− B/2
(3.3.2)

Where additionally w(y) is the downwash distribution on the wing.

The LS(1)-0417MOD airfoil model has a rectangular wing platform, which simplifies

solution of the equations with a elliptical lift distribution, which is defined as follows,

Γ ( y) = Γ 0 1 − ( )
2y 2
B (3.3.3)

This simplification allows for the integration of the equations and the resultant lift and drag

are

L = ρVΓ 0 Bπ / 4
(3.3.4)
D = ρΓ 02 π / 8

36
The constant '0 is a scaling factor for each individual case. In this case the lift coefficient

was nondimensionalized to the value of 1, and then the induced drag coefficient becomes,

CL 2
C Di = (3.3.5)
πeAR

Where e is the span efficiency factor defined as follows

1
e= (3.3.6)
1+ δ

Where * is the correction factor for induced drag and for a nontapered wing it is 0.05,

therefore e equals to 0.952 [47].

1.5

Clean
Theory
k/c = 0.0019
Cl max Clean

1.0
Lift Coefficient

0.5

For α = 10°

Steady State
Re = 1.0 million

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Span, z

Figure 23. 3D theory comparison.

37
The previously discussed theory was applied to the studied model and the obtained

results and experimental clean and grit applied data are plotted in figure 23. The agreement

in this example for 10/ angle of attack, between the clean case and the theory is very good,

but for the grit applied case the agreement departs slightly, since the theory does not take into

account the turbulence produced by the grit particles.

The complexity and difficulty in obtaining 3D steady state and unsteady pressure
distributions and aerodynamic characteristics were proven by a blind test for the National
Renewable Energy Laboratory. The results from those tests showed that the individual codes
can give 60% under-prediction or 150% over-prediction in calculation of the turbine torque
[36]. In many of those cases experimental data are still needed to calibrate and verify these
codes.

38
CHAPTER 4

RESULTS AND DISCUSSION

4.1 Steady State Data: Clean and Roughness Applied with Vortex Generators

-3

Clean 2D data
z=0.925
Re=1 million z=0.742
Cl=0.5 z=0.592
-2 z=0.408
z=0.258
z=0.075
Pressure Coefficient

-1

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

X/C

Figure 24. Cp distribution for Cl=0.5, clean.

The steady state data presented in this section are for both 2D and 3D model
configurations. These steady state tests provide an essential baseline for comparison with

39
the vortex generators in steady state flow and will be later used in comparison to the unsteady
flow data, where the flow field changes. The clean data are used as a baseline for
comparison with the leading edge grit roughness and vortex generators cases. The pressure
coefficients for all the conditions were acquired for all six stations across the span of the
model, giving a good flow representation over the airfoil surface. The differences in the flow
characteristics between clean, grit applied or vortex generators are visible from the pressure
distributions. These details can explain some of the flow changes due to the application of
grit or vortex generators. The pressure coefficients were later integrated to determine the lift,
pitching moment and pressure drag coefficients.
The pressure distributions reveal many flow characteristics such as separation, 3D effects
and changes due to application of grit. In figure 24, the pressure distributions at all span
stations including 2D and for the baseline clean model are shown for Reynolds number of
1 million and a lift coefficient of 0.5. The pressure distributions were interpolated to an
equal lift coefficient value, so that they could be easily compared, since each individual angle
of attack may have been at a slightly different value. There is little difference between 2D
and 3D pressure distributions at this low lift coefficient of 0.5. Only the tip station shows
a slightly lower pressure on the last 20% of the airfoil and a slightly higher one around the
leading edge. The pressure peak at approximately 10% of the chord was -1.25 for all span
locations and 2D data.
The assumption used to develop 3D theory is proven here in the similarity of the pressure
distributions of the 2D cases and 3D cases at a particular lift coefficient. The angle of attack
for each of those pressure distributions will actually be different due to the different lift curve
slope for the stations. Also, interactions with the tip vortex and individual station stall
characteristics at higher angles of attack limit this comparison to approximately 5/ less than
the stall angle of attack.

40
-3

Clean 2D data
z=0.925
Re=1 million z=0.742
Cl=1.15 z=0.592
-2 z=0.408
z=0.258
z=0.075
Pressure Coefficient

-1

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

X/C

Figure 25. Cp distribution for Cl=1.15, clean.

The pattern at the tip station (z=0.925) becomes more explicit at the next lift coefficient
value of 1.15, as shown in figure 25. In this case the pressure peak reaches approximately
-2.5 for all the sections, except the tip where it is -2.25. The change in the pressure
distribution at the tip is due to the spanwise flow issuing from the tip vortex. The flow
differs from 2D in that there is a cross-flow component, the flow is now curling over the tip
of the model from the bottom side to the top for the positive angle of attack. For higher lift
coefficient values, this effect will eventually be seen in the next span section, that will occur
closer to stall. As can be seen the span of the influence of the tip vortex is limited to only
the one tapline. Already the 0.742 station and 2D data are very close together. The
assumption made when dealing with 3D data that the pressure distribution for a 3D wing is
the same as the 2D at the same lift coefficient value, holds true for all the stations, except at
the tip, in the linear region.

41
-3
Flat Vgs 2D data
z=0.925
Re=1 million z=0.742
Cl=1.0 z=0.592
z=0.408
z=0.258
-2 z=0.075
Pressure Coefficient

-1

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

X/C

Figure 26. Cp distribution for Cl=1.0, Flat vortex generators.

The application of the vortex generators does not change the shape of the pressure
distributions in comparison to the clean configuration in the linear region, such as lift
coefficient up to 1.15, as shown in figure 26. The same slight differences in the individual
span location are repeated for the vortex generators cases, for example the 2D data has a
slightly lower peak value and has a lower pressure coefficient from 10% to 70% on the upper
surface and from 0% to 20% on the lower surface. This is the only departure of the 2D data
from 3D, excluding the tip section. The tip station is still the only pressure distribution
which shows the effects of the tip vortex. The flow is not stalled on the section because the
pressure is not constant (perfectly flat), stall responses can be seen later in the grit applied
cases.

42
-4
Flat Vgs 2D data
z=0.925
Re=1 million z=0.742
Cl=1.5 z=0.592
-3 z=0.408
z=0.258
z=0.075
Pressure Coefficient

-2

-1

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

X/C

Figure 27. Cp distribution for Cl=1.5, Flat vortex generators.

The pressure distributions with Flat vortex generators applied are shown in figure 27, for
a lift coefficient of 1.5 and Reynolds number of 1 million. The 2D data still has the same
shape of the pressure distribution with a peak value of -3.4. The tip station at this high lift
shows a much lower pressure peak at -3.0 and a lower pressure at the trailing edge. Where
as, the rest of the data peaks between -3.75 to -4.0. There is more variation at this lift
coefficient value in the distributions between the sections, especially at the tip where the
effect of the tip vortex is becoming more noticeable. The clean data is not available for lift
coefficient of 1.5 because without the vortex generators this lift is not achieved only by the
clean wall section and that particular section does not show tip vortex effects.

43
2.1

1.8

1.5
Lift Coefficient

1.2

0.9
Clean
Flat
Curled
0.6 Type B
Type A

2D Data
0.3 Steady State

Vortex generators were positioned at 0.1 of the chord

0.0
0 10 20 30

Angle of Attack, deg

Figure 28. 2D lift coefficient.

The pressure distributions such as those previously shown were integrated to obtain lift,
pitching moment and pressure drag coefficients. The 2D steady state lift coefficient plot is
shown in figure 28. Even through the vortex generators were not optimized for the 2D runs,
the increase in the lift coefficient was observed. The position of the vortex generators was
selected as an average optimal place for the 3D testing (0.1 chord). The lift curve slope for
all of the 2D, clean with or without vortex generators, was calculated at 0.115. To increase
the lift coefficient more the vortex generators should be positioned further down the chord
as found in other tests. The vortex generators did maintain a linear lift curve slope until stall
whereas, by comparison the clean data show a departure at 10/ angle of attack, approximately
5/ before stall. Type A vortex generators did not do very well, they generated lower lift
coefficient than even 2D clean data. The ineffectiveness of the Type A vortex generators is
due to their position and a relatively thick base (0.05"). That high base prevented the vortex

44
generators from being in the boundary layer at the testing point. Moving them to a further
aft location would improve their performance, as shown in other tests. The best lift
coefficient is obtained with the Flat vortex generators and it is 1.77 at 14.1/ angle of attack,
in comparison with the clean maximum lift coefficient which is 1.58 at 14.4/. The vortex
generators increased the lift by 12% and with the vortex generators the stall was mild and the
lift was higher than the clean 2D case for more than 10/. Additional data can be found in
Appendix B table B1.

2.1

1.8

1.5
Lift Coefficient

1.2

0.9
Clean
Flat
Curled
0.6 Type B
Type A

2D Data
0.3 Steady State

Vortex generators were positioned at 0.1 of the chord

0.0
0.00 0.01 0.02 0.03 0.04 0.05

Wake Drag Coefficient

Figure 29. 2D drag polar.

The pressure drag was not used as a reference since it is not accurate at low angles of
attack, because there are not enough pressure taps around the leading edge of the airfoil and
the pressures there cannot be resolved. The pressure drag can be used after stall where the
drag coefficient is high and where the wake is too wide and erratic and cannot be acquired

45
by the probe, therefore the data from it is not accurate. The wake drag was measured to
obtain accurate drag coefficients for the 2D configuration that would be used to calculate the
drag for 3D configuration. The drag polar for the 2D data and the different vortex generators
is shown in figure 29. The application of the vortex generators increases the drag coefficient
in comparison with the clean data. The Type B vortex generators were only 0.15" high and
increased the drag only slightly, about 5%. The Flat and Curled vortex generators increased
the drag significantly due mostly to their lager size. The Flat vortex generators increased the
drag by 25% over the clean baseline and the Curled by 80%. The minimum clean wake drag
coefficient was approximately 0.0100 and with the Flat vortex generators the drag increased
to 0.0150. The Flat and Type B vortex generators decrease the drag coefficient above a lift
coefficient value of 1.4, which in the clean case corresponds to the onset of severe stall on
the upper surface of the model.

2.1

1.8

1.5
Lift Coefficient

1.2

0.9

Clean
0.6 Type B at 0.07 chord
Type B at 0.15 chord
Steady State
Flat Cut
z=0.075 Flat
0.3 Curled

Vgs slanted from tip to root from 0.07 to 0.15 chord


0.0
0 10 20 30

Angle of Attack, deg

Figure 30. Lift coefficient at z=0.075.

46
Figure 30 shows the lift coefficient at the wall station (z=0.075) for clean model and with
the vortex generators applied. The Type B vortex generator data shown here include two
positions one at 0.07 chord and at 0.15. They are shown for comparison of the
unpredictability of the vortex generators. The 0.07 position did not increase the lift
coefficient at all, it even lowered the maximum lift coefficient from the baseline by 0.05.
The 0.15 chord position increased lift to 1.6 and the stall point was moved by 3/ to 20/ angle
of attack. This behavior suggested that the vortex generators at the wall have to be
positioned further aft to affect the lift coefficient. The Flat and Curled vortex generators
were applied at the slant from 0.07 chord at the tip to 0.15 chord at the wall, since previous
tests showed this as an ideal position fo the 3D testing. This set up was used to obtain all the
3D data with the vortex generators. The application of a certain type of the vortex generators
increased the maximum lift coefficient significantly. The lift curve slope was calculated to
be 0.081 for all the presented data. It was shifted by 0.05 due to a slight decambering effect
caused by the application of the vortex generators. As can be seen from the figure, the Flat
and Curled vortex generators increased the lift coefficient by approximately 26%. The clean
baseline maximum lift coefficient of 1.52 was obtained at 17.4/, while with the vortex
generators stall was delayed to 24/ angle of attack. The maximum lift coefficient was 1.92
for the Flat vortex generators at 23.4/, the Curled gave a slightly higher lift coefficient. The
3D steady state data are summarized in Appendix B table B6 and B7.

47
Clean
Type B at 0.07 chord
Type B at 0.15 chord
Flat Cut
Steady State Flat
Curled
0.0 z=0.075
Vgs slanted from tip to root from 0.07 to 0.15 chord
Pitching Moment Coefficient

-0.1

-0.2

-0.3
0 10 20 30

Angle of Attack, deg

Figure 31. Pitching moment coefficient at z=0.075

The pitching moment coefficients for clean and vortex generators applied conditions at
the wall station are shown in figure 31. The pitching moment is not affected by the
application of the vortex generators in the linear region and does not show the shift as the lift
curve did. It remained almost constant at -0.075 from 0/ angle of attack to stall for both the
clean data and the sets with the vortex generators. The pitching moment drops close to the
stall angle of attack by 0.05 for the Flat and Curled vortex generators and to almost 0.1 for
the Type B, due to earlier and more sever stall. In comparison, the clean data close to stall
drops close to 0.05. At high angles of attack, above 25/, the Flat, Curled and Type B at 0.15
chord all fall in together with the clean baseline.

48
2.1

1.8

1.5
Lift Coefficient

1.2

0.9
Clean
Type B at 0.15 chord
Type B at 0.07 chord
0.6 Flat-Cut
Flat
Curled

0.3 Steady State


z=0.925
Vgs slanted from tip to root from 0.07 to 0.15 chord
0.0
0 10 20 30

Angle of Attack, deg

Figure 32. Lift coefficient at z=0.925

At the tip station of z=0.925, the vortex generators had a lesser effect on the lift
coefficient than at the wall station, as can be seen from figure 32. As before the clean
baseline data are plotted together with the three sets of vortex generators. Type B vortex
generators were also run at two positions, 0.07 and 0.15. For the tip station there is a slight
increase in lift when the vortex generators are positioned at 0.07 chord in comparison to 0.15.
The lift increases from 1.38 to 1.40 and the maximum lift occurs at the same angle of attack
of 22/ for both positions. The maximum lift coefficient was 1.2 at 18.3/ for the clean model
and the lift curve slope was 0.0542 which is less than that at the wall section, which has a
slope of 0.081. When all the sections are plotted together the characteristic fan of the data
occurs with the wall and tip stations serving as boundaries, with highest and lowest slopes.
The Flat and Curled vortex generators did increase the lift coefficient significantly but not
as high as at the wall station where the effect of the tip vortex was much lower. The

49
maximum lift coefficient obtained was 1.6 by both the Flat and Curled vortex generators at
24/ angle of attack, which is 33% higher than the clean baseline model. The lift curve slope
with the application of the vortex generators exhibits at 8/ angle of attack a change in slope
from 0.047 up to 8/ to 0.069 through the rest of the linear region. That change of slope can
be attributed to the interaction of the vortex generators with the tip vortex.

Vgs slanted from tip to root from 0.07 to 0.15 chord


Clean
Type B at 0.15 chord
Type B at 0.07 chord
Flat-Cut
0.0 Flat
Curled
Pitching Moment Coefficient

Steady State
z=0.925

-0.1

-0.2

-0.3
0 10 20 30

Angle of Attack, deg

Figure 33. Pitching moment at z=0.925

The tip station quarter chord pitching moment with the applied vortex generators is
shown in figure 33. The pitching moment is no longer linear at the tip station, a parabolic
curve could be fitted through the data. All the individual plots have a similar shape to that
of the clean pitching moment, but the vortex generators applied case show a substantial
decrease for angles of attack higher than 8/, where the lift curve slope changed. At this

50
particular angle of attack the interactions of the vortex generators with the boundary layer
and the tip vortex must change. At the stall angle of attack of 24/, the value of the pitching
moment for the Flat and Curled vortex generators is -0.25, in comparison the clean baseline
has a value of -0.13; this is a 10% decrease. This station does not show a pitching moment
drop after stall, as does the wall station. In this case the moment increases slightly after stall,
this is attributed to the strong effects of the tip vortex.

-3

Grit 2D data
z=0.925
Re=1 million z=0.742
Cl=1.0 z=0.592
-2 z=0.408
z=0.258
z=0.075
Pressure Coefficient

-1

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

X/C

Figure 34. Cp distribution for Cl = 1.0 with grit

The application of the leading edge grit roughness simulated ageing of the model or an
accumulation of insects or dust. As can be seen from figure 34, the application of the leading
edge grit roughness changes significantly the pressure distributions, especially close to the
tip of the wing. This figure shows the pressure distribution with the grit applied for all the
spanwise stations and the baseline 2D data at a lift coefficient of 1.0, except for the tip

51
station, which stalls at a lift coefficient of 0.82. In this case the tip station at z=0.925 does
not exhibit spanwise flow from the vortex but separation from 0.05 of the chord to the
trailing edge. The roughness exhibited in the pressure distribution is attributed to the grit
particles being close to the tap and unsteady stall characteristics. The separated flow extends
through 85% of the chord at the 0.742 spanwise station and the last 10% of the chord for the
0.592 station pressure distribution. The other pressure distribution have the same shape as
the clean or vortex generators applied data at the same lift coefficient value. For the lower
lift coefficient value of 0.5 all the pressure distributions look exactly like the clean cases for
the same lift, as seen before. Stall does not occur and the cross-flow is only limited to 20-
30% of the tip station. The angle of attack for the individual pressure distribution plots will
be different since the lift depends on the model condition, for example grit applied or vortex
generators. For the lift coefficient of 1.15 the data could not be plotted since stall occurs
earlier than 1.15 for all the section but the 2D and the wall station (z=0.075).

52
-3

Grit + Flat 2D data


z=0.925
Re=1 million z=0.742
Cl=1.15 z=0.592
-2 z=0.408
z=0.258
z=0.075
Pressure Coefficient

-1

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

X/C

Figure 35. Cp distribution for Cl=1.15, grit and VGS.

The application of the Flat vortex generators “restores” the pressure distribution back to
the baseline characteristics as can be seen from figure 35, which shows the leading edge grit
roughness applied and the Flat vortex generators for lift coefficient value of 1.15. The slight
“roughness” of the pressure distribution at the leading edge is due to the grit strip and the
particles being close to the taps. For the lower lift coefficient value of 0.5 the pressure
distribution exhibit a slight flat at the peak, due to the grit strip and vortex generators
positions. Also the cross-flow in the tip section is more visible, it occurs for about 30% of
the chord. The application of the vortex generators does increase the lift coefficient but the
stall angle of attack is still lower than that of the clean data.

53
2.1

1.8

1.5
Lift Coefficient

1.2

0.9

0.6

Steady State No VGs


z=0.075 Flat.
0.3 Curled
LEGR applied
Vgs slanted from tip to root from 0.07 to 0.15 chord
0.0
0 5 10 15 20 25

Angle of Attack, deg

Figure 36. Lift coefficient at z=0.075, grit with VGS

Figure 36 shows the lift coefficient for the wall section with the LEGR cases with and
without vortex generators. The application of the grit has lowered the maximum lift
coefficient by 24% from 1.52 at 18/ angle of attack for the clean model to 1.16 at 13/. The
leading edge grit roughness also changes the stall angle of attack by 28%. The maximum lift
coefficient with the clean vortex generators was 2.0 at 24/ and for the vortex generators with
the grit applied gave a lift of 1.49 at 18/. The lift coefficient decreased 26% and the stall
angle of attack also decreased by 25%, these percentages are very similar to that of the clean
and grit only. The application of the vortex generators to the grit strip increases the lift
coefficient and extends the linear range of the curve. The maximum lift coefficient for the
vortex generators with the LEGR was 1.65 for the Curled and 1.49 for the Flat. The
maximum lift coefficient for the grit was 1.16 and 1.49 with the Flat vortex generators as
mentioned earlier. This corresponds to a 28% increase for the Flat and a 42% increase for

54
the Curled. The stall angle of attack changed from 13/ with no vortex generators to between
18/ and 20/ for the cases with the vortex generators. In this case, the Curled vortex
generators were slightly less sensitive to the application of the LEGR than the Flat ones.

0.0
Pitching Moment Coefficient

-0.1

-0.2 Steady State No VGs


z=0.075 Flat
Curled
LEGR applied

Vgs slanted from tip to root from 0.07 to 0.15 chord

-0.3
0 5 10 15 20 25

Angle of Attack, deg

Figure 37. Pitching moment at z=0.075, grit with VGS

The quarter chord pitching moment at the wall station for both the model with and
without the vortex generators applied is shown in figure 37. It was almost not affected by
the application of the LEGR. The application of the vortex generators also does not change
the pitching moment which is found to be approximately -0.075 until stall, where it drops to
0.05. In comparison without the vortex generators the moment is -0.06, this is a slight
change from the clean moment coefficient which was -0.075, as the vortex generators data.

55
2.1

1.8

1.5
Lift Coefficient

1.2

0.9

No VGs
0.6 Flat
Curled
Steady State
0.3 z=0.925
LEGR applied
Vgs slanted from tip to root from 0.07 to 0.15 chord
0.0
0 5 10 15 20 25

Angle of Attack, deg

Figure 38. Lift coefficient at z=0.925, grit with VGS

Figure 38 shows the lift coefficient for the tip section (z=0.925) with the LEGR applied
for configurations with and without vortex generators. The maximum lift coefficient with
applied grit only dropped to 0.82, a 32% decrease in comparison to the clean airfoil, which
had the maximum lift coefficient of 1.20. The stall angle of attack changed from 18/ for the
clean to 13/ for the grit. With the vortex generators applied the lift coefficient increased to
1.15, a 40% increase from the grit only case. Both sets of vortex generators, the Flat and
Curled, performed almost exactly the same. The difference in slope of the two curves at
angles of attack less then 8/ can be compared to the clean vortex generators data where the
same slope change occurred. The stall angle of attack has moved to 15/ from 24/ when the
model was clean. The increase of lift is still significant with respect to the baseline model
with the grit applied. The clean vortex generators obtained a lift coefficient of 1.6 at 24/, so

56
with respect to the clean vortex generators model case the application of this roughness
produces a 28% drop in lift and a 25% decreases in the stall angle of attack.

0.0
Pitching Moment Coefficient

-0.1

-0.2
Steady State No VGs
z=0.925 Flat
Curled
LEGR applied
Vgs slanted from tip to root from 0.07 to 0.15 chord
-0.3
0 5 10 15 20 25
Angle of Attack, deg

Figure 39. Pitching moment at z=0.925, grit with VGS

In figure 39 the tip station quarter chord pitching moment is shown for the LEGR applied
model condition with and without vortex generators. The moment coefficient is less negative
with the grit applied for the whole angle of attack range tested. At 0/ angle of attack the
pitching moment is only slightly more positive than the clean case, it stays constant at that
value of -0.05 up to approximately 6/ angle of attack. The moment then decreases almost
linearly to -0.15 at 25/angle of attack. The application of the vortex generators changes the
pitching moment shape back to what the clean data looked like. The actual minimum of the
pitching moment with the vortex generators applied falls in the range of the clean data.

57
4.1.1 Steady State Integrated Data

2.1

1.8

2D data
1.5
Lift Coefficient

1.2

0.9 z=0.925
z=0.075

0.6 Symbols integrated lift coefficient

0.3 Steady State


Clean

0.0
0 10 20 30

Angle of Attack, deg

Figure 40. Integrated lift coefficient, clean

Each of the six pressure distributions was integrated to obtain local lift, moment and
pressure drag coefficients, then the total lift, moment and pressure drag coefficients were
obtained by integrating the local coefficients across the span for each angle of attack. The
data were not acquired exactly at z=0 (the tip) and z=1 (the wall), so the z=0.075 and
z=0.925 points were used at z=0 and z=1 for full integration, respectively. Figure 40 show
the clean baseline total lift coefficients. For comparison 2D data are shown additionally with
the tip and wall station lift curves. The integrated data fall between the tip and wall stations.
The local lift was integrated up to stall only, since after the stall the data are less reliable due
to unsteady stall characteristics. The data after stall could not be always duplicated exactly,
because the flow characteristics there depend on the behavior of the boundary layer and the

58
exactly same angle of attack could not be guaranteed with the present experimental setup for
steady state runs.

2.1

1.8

2D data

1.5
Lift Coefficient

1.2

z=0.925
0.9
z=0.075

Symbols integrated lift coefficient


0.6

Steady State
0.3
Flat Vgs

0.0
0 10 20 30
Angle of Attack, deg

Figure 41. Integrated lift coefficient, Flat VGS

For 1 million Reynolds number the dependence of the lift on the angle of attack is linear
up to stall of 17/angle of attack, shown in figure 41. The integrated data had a slope of
0.074, in comparison to that of the 2D clean data slope of 0.118. The maximum total lift
coefficient for the clean configuration was found to be approximately 1.4 at 17/. The
integrated 3D curve falls between the wall and tip station data, usually closer to the wall
station plot. This is true for most of the integrated data, since usually only the tip station
showed strong cross-flow effect, which would lower the slope of the lift coefficient curve.
The slight shift of the integrated lift plot at about 10/ is attributed to data scatter.

59
Figure 41 also shows the integrated data but the Flat vortex generators were applied to
the model surface in this case. As can be seen from the figure, there is less of a slope change
between the tip and the wall stations due to the suppression of stall and control of the
cross-flow offered by the vortex generators. In this case the integrated lift curve lies almost
exactly half way between the tip and wall stations. In the low angle of attack range the plot
falls on the tip station, which is unusual, because in all the other cases it follows the wall
station at low lift coefficient values. This new trend suggests that the local lift obtained with
the vortex generators at the stations closer to the wall is much lower than for the clean model.
The average slope of the integrated data was found to be 0.075 in comparison the vortex
generators tip station obtained 0.0712 and the wall station 0.089. The average slope in this
case is closer to that of the tip station. The maximum total lift coefficient was 1.82 at 24/
angle of attack. The application of the Flat vortex generators gave a 30% increase in total
lift coefficient with respect to the clean model.

2.1

1.8

1.5
Lift Coefficient

1.2
2D data

0.9

z=0.925
0.6 z=0.075

Symbols are integrated lift coefficient

0.3 Steady State


k/c = 0.0019

0.0
0 10 20 30

Angle of Attack, deg

Figure 42. Integrated lift coefficient, grit

60
The application of the grit lowered the lift curve slope of the integrated data in
comparison with the clean data, as shown in figure 42. The lift curve slope is 0.056 for the
integrated data with grit applied, the clean case had a value of 0.074. It is a 24% decrease
in slope due to the application of the grit strip. The 2D slope was 0.12, much higher than
with the grit applied. As mentioned before, the integrated data lie closer to the wall section
than the tip. The maximum lift coefficient was found to be 1.0 at 14/ angle of attack. The
lift coefficient decreased 40% due to the application of the grit, when the data is compared
with the baseline clean model.

2.1

1.8

1.5 2D data
Lift Coefficient

1.2

0.9 z=0.925
z=0.075

0.6 Symbols integrated lift coefficient

Steady State
0.3 k/c=0.0019
Flat Vgs

0.0
0 10 20 30

Angle of Attack, deg

Figure 43. Integrated lift coefficient, grit with VGS

As seen from figure 43, the application of the vortex generators together with the grit
increased the total lift coefficient up to 1.4 found at 18/. That is a 40% increase in lift due
to the vortex generators when compared to the gritted model condition. The lift is still much

61
lower than for the clean model with the vortex generators, which had a total lift coefficient
of 1.82 at 24/ angle of attack. In the linear region the slopes of the grit vortex generators data
and the clean vortex generators data are exactly the same, but past 4/ angle of attack the dirty
model slope decreases. The average slope was calculated to be 0.072. Stall, of course,
occurred earlier with the grit since it introduces turbulence to the boundary layer and makes
it less stable. In comparison to the grit only data, the vortex generators on average did not
change the slope of the total lift curve slope, but delayed stall and therefore, increased lift.

g g
20

10
Angle of Attack, deg

C =C +C
D dw Di
z=0.925
0 z=0.742
z=0.592
z=0.408
z=0.258
z=0.075

Clean - 3D model configuration


Symbols intergrated induced drag coefficient

-10
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10

Drag Coefficient

Figure 44. Total drag polar, clean

Figure 44 shows the total drag coefficient for the clean model, defined as a summation
of the viscous drag and the induced drag. The viscous drag for the 3D case data was
estimated using the actual pressure distributions and a calibrated boundary layer code. The
calibration of the code was done by using 2D pressure distributions and the wake data taken.

62
The code has two sets of multipliers which allow changes to the turbulence levels in both the
laminar and turbulent parts of the code. These multipliers were adjusted to obtain the correct
wake drag for the 2D data and then used with the 3D pressure distributions. The local
induced drag was calculated by multiplying the local lift coefficient and the local induced
angle of attack. The individual induced drag coefficients were calculated for 1 million
Reynolds number by using both the 2D data and data from the six 3D stations. The induced
angle of attack can be defined as a difference between the flow direction of the local 3D flow
and 2D flow. Using that idea, for each angle of attack in 3D, the lift was found on a plot and
the corresponding 2D angle of attack was then determined from the 2D lift coefficient verses
alpha data. That difference is the induced angle of attack The calculated viscous drag varied
for the different stations. In the low angle of attack range where the viscous drag is a large
part of the total drag, it was found to be approximately 0.0120 for all the stations and for
simplicity one average value was used. The induced angle of attack was found to be zero at
-4/ angle of attack. The total drag is shown in open symbols in figure 44. The total drag
coefficient increases parabolically from a minimum of 0.012 and for example at 10/ angle
of attack the total drag is already 0.09.

63
1.5

1.0
Lift Coefficient

Clean
0.5 Flat Vgs
k/c=0.0019
Flat Vgs and Grit

C =C +C
D dw Di

0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10

Drag Coefficient

Figure 45. Total drag polar

Figure 45 shows the total drag coefficient without all the details for all the cases
described before. As can be seen from the figure, the lowest total drag is obtained by the
clean model condition, where the minimum total drag coefficient was 0.012. The application
of the grit increased the drag coefficient at the lift coefficient values of less than 0.5, but at
higher lift the vortex generators actually gave a slightly lower total drag value. The Flat
vortex generators did not increase the drag coefficient at low angle of attack, only at lift
coefficient value of 0.4, the increase was noticeable. The Flat vortex generators and grit
together generated the highest increase in drag.

64
4.2 Unsteady Data

The pitch oscillations can simulate unsteady incoming flow or gusting and provide
information on the airfoil characteristics response to these phenomena. The understanding
of these flow characteristics is necessary in order to provide more detailed information which
can be introduced into the unsteady flow CFD. The main difference between the unsteady
and steady state flows is the development of the hysteresis loops due to a lagging boundary
layer. The boundary layer in unsteady flow is unable to keep up with the angle of attack
change, therefore a thinner boundary layer is formed at increasing angles of attack before
stall. This gives a higher lift coefficient value before the boundary layer separates from the
surface but after stall the boundary layer becomes much thicker and the lift therefore smaller.
The lift coefficient can increase up to 70% depending on conditions, this load change has to
be considered in structural design to prevent failure of the wing.
Unsteady 2D experimental data were obtained for the LS(1)-0417MOD 3D airfoil model
undergoing sinusoidal pitch oscillations. As mentioned earlier, no calibration for the tunnel
contraction pressures was available for the unsteady oscillating model conditions and the
steady state tunnel calibration was used to set the flow conditions while the model was
stationary at its mean angle of attack. A comprehensive set of test conditions was used to
describe unsteady behavior of the airfoil including: two angle of attack amplitudes, ±5.5/ and
±10/; three Reynolds numbers, 0.75, 1, and 1.25 million; three pitch oscillation frequencies,
0.6, 1.3, and 1.8; and three mean angles of attack, 8/, 14/, and 20/. The extensive test matrix
provided much more data than could be presented here, so in order to show trends in the data,
only results from 1 million Reynolds number are shown with the low and high oscillating
frequencies, for the high amplitude case of ±10/.

4.2.1 2D Unsteady Data

A typical example of clean lift coefficient behavior at a 1.8 Hz oscillating frequency for
the three different mean angles is shown in figure 46. It also includes the 2D steady state

65
data for comparison. As can be seen from the figure, the lift coefficient increases to
approximately 2.3, which is 53% higher than the clean steady state, which had a maximum
lift coefficient of 1.55. The high increase in the lift coefficients due to unsteady flow
conditions leads to a significant increase in loading, which can lead to structural fatigue or
failure when not included in the design. The 20/ mean angle case reaches the same lift
coefficient as the 14/ mean angle, in this particular case. Usually the 20/ case still increases
the lift coefficient slightly and has sharper stall characteristics. The difference between the
upswing and downswing lift coefficient values at the same angle of attack defines a measure
of the hysteresis loop. The loops become more pronounced at the higher mean angle. The
hysteresis loops occur because, in this type of flow, the boundary layer does not always have
enough time to catch up with the outer flow. Especially after stall the boundary layer is
usually still separated even when the angle of attack is already low and on the upswing the
flow is attached, since a change in the displacement thickness has occurred due to the history
of the boundary layer. At higher frequencies and mean angles the boundary layer is different
than on the upswing until low angle of attack, only close to 5/ the lift coefficients become
equal between the upswing and downswing.

66
2.5

2.0
Lift Coefficient

1.5

1.0

o
0.5 8 omean angle Clean, 2D
14o mean angle
20 mean angle
Re=1 million,
Steady state f = 1.8 Hz, ± 10° Sin
0.0
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 46. 2D unsteady lift coefficient, clean, f = 1.8H z.

To measure the size of a hysteresis loop two lift coefficient values at the same angle of
attack were selected, one on the upswing and the other on the downswing. For example for
8/ mean angle at 12/ angle of attack, the upswing lift coefficient was 1.55 and the
downswing was 1.00. The maximum lift coefficient of 2.3 occurred at 22/ and the
corresponding hysteresis lift was 1.00. The change in lift coefficient at 8/ mean angle was
0.55, a 35% drop in lift in comparison to the upswing. Meanwhile at 14/or 20/ mean angle
cases (in this case both had the same maximum lift coefficient), the upswing lift difference
was 1.3, a 57% drop. These changes between the upswing and downswing are significant.
The loads n the actual airfoil change drastically, these changes have to be incorporated into
the structural design, since otherwise structural failure may occur. The 2D oscillating data
are summarized in Appendix B, tables B2 to B5.

67
0.1

Pitching Moment Coefficient -0.0

-0.1

-0.2

-0.3

o
8 omean angle Clean, 2D
-0.4 14o mean angle
20 mean angle
Re=1 million,
Steady state f = 1.8 Hz, ± 10° Sin
-0.5
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 47. 2D unsteady pitching moment coefficient, clean, f=1.8 Hz.

The quarter chord pitching moment coefficient for the same case as the lift is shown in
figure 47. The moment coefficient is not affected by the unsteady behavior as much as the
lift. The very large hysteresis loops occur at high angles of attack, above 20/. The steady
state pitching moment coefficient is almost constant at -0.1 until 10/ angle of attack, at that
point the 8/ mean angle case has a upswing quarter chord pitching moment value of -0.11
and a downswing value of -0.01. The pitching moment generally for upswing is more
negative than the downswing. Usually the steady state pitching moment data fall close to the
upswing pitching moment and the downswing values are closer to 0. For angles of attack
higher than 20/ the steady state pitching moment coefficient tracks that of the downswing.
For example, the pitching moment at 22/ angle of attack is -0.4 on the upswing and -0.11 on
the downswing. The change from steady state value at that point is 26% for the upswing and
the downswing exhibits the same values as steady state.

68
2.5

2.0
Lift Coefficient

1.5

1.0

o
0.5 8 omean angle Clean, 2D
14o mean angle
20 mean angle
Re=1 million,
Steady state f = 0.6 Hz, ± 10° Sin
0.0
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 48. 2D unsteady lift coefficient, clean, f = 0.6 Hz

Figure 48 represents the 2D unsteady lift coefficient for the low frequency, 0.6 Hz. As
can be seen from the figure the effect of pitch oscillations is greatly reduced for low
frequency in comparison to 1.8 Hz. The hysteresis loops are approximately half of the size
of those for the high frequency case. The upswing and downswing lift for the low mean
angle of 8/ is almost the same from 0/ to 8/ angle of attack, in comparison the lift was equal
only up to 4/ for the 1.8 Hz frequency. The hysteresis loop that occurs close to stall is
relatively small with the upswing lift coefficient of 1.6 at 12/ angle of attack and the
downswing lift of 1.3. The lift coefficient change was only 0.3, whereas for the high
frequency the change was 0.55. The hysteresis loop size was approximately doubled with
the frequency increase from 0.6 to 1.8 Hz. The maximum lift coefficient at 8/ mean angle
is 1.8 and on the downswing the lift is 1.5. The hysteresis loop increases in size for the 14/
mean angle and then decreases for the 20/, since at those high angles of attack for the low

69
frequency the airfoil surface is largely stalled. The 14/ and 20/ mean angles produce the
same shapes of the lift coefficient curves only shifted in angle of attack. The maximum lift
coefficient for the 0.6 Hz case was 1.86 at 18/ and the hysteresis lift coefficient was 1.15, this
represents a 38% drop in comparison to the upswing. While for the 1.8 Hz case the
maximum lift coefficient was approximately 20% higher than for the 0.6 Hz.
The quarter chord pitching moment coefficient has a similar trend to the lift coefficient
at the low frequency, in that the hysteresis loops are smaller. The moment coefficient at 12/
angle of attack for the upswing is -0.08 and the hysteresis moment was found to be equal to
that of steady state at that angle of attack (-0.06). Since the hysteresis loops are much
smaller, the pitching moment coefficient is closer to that of steady state. At 20/ angle of
attack, where stall occurred, the pitching moment was on the downswing equal to the steady
state value of -0.1. The hysteresis moment coefficient was between -0.12 to -0.15, due to
less stable boundary layer response to the oscillations at the low frequency. Generally more
data scatter was observed at the low frequency and some of the scatter can be attributed to
the experimental setup.

70
2.5

2.0
Lift Coefficient

1.5

1.0

0.5 Grit, 2D
o
8 omean angle
14 mean angle
Re=1 million,
Steady state f = 1.8 Hz, ± 10° Sin
0.0
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 49. 2D unsteady lift coefficient, LEGR, 1.8 Hz.

The application of the leading edge grit roughness (LEGR or grit), which changes the
boundary layer thickness and its behavior, affects the overall flow characteristics
significantly. As can be seen from figure 49, the steady state maximum lift coefficient is 1.1
and the increase due to the pitch oscillation is 100%. The maximum lift coefficient was
found to be 2.1 at 18/ for the 14/ mean angle case. The hysteresis lift was 0.8, that is a 62%
drop from the maximum. In comparison the clean data had a maximum lift coefficient of 2.3
at 22/ angle of attack. The 20/ mean angle pressure distributions showed general stall at
angle of attack higher than 15/ and the lift obtained with these oscillations was lower than
the 14/ mean angle cases. This phenomena occurs in experimental data where severe leading
edge grit roughness was applied, because the boundary layer in these high mean angle
oscillations is thick and unstable. It is also difficult to predict the response of the model at
the higher mean angles (well after the linear region of the lift curve) to the grit or vortex

71
generators. The size of the hysteresis loops has increased greatly and they extend all the way
to 2/ angle of attack, especially at 8/ mean angle. The lift at 12/ angle of attack was 1.50 and
the downswing value was 0.6, this creates a 60% drop. In comparison the clean lift at that
point was only 1.55 but on the downswing the lift coefficient dropped only 35% to 1.00.

0.1
Pitching Moment Coefficient

-0.0

-0.1

-0.2

-0.3

o
8 omean angle Grit, 2D
-0.4
14 mean angle Re=1 million,
Steady state
f = 1.8 Hz, ± 10° Sin
-0.5
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 50. 2D unsteady pitching moment coefficient, LEGR, f = 1.8 Hz.

The quarter chord pitching moment coefficients for the same conditions are shown in
figure 50. The moment coefficient is very close to the clean case until stall occurs, from that
angle of attack the moment is more negative and large hysteresis loops occur at the 14/ mean
angle. The pitching moment becomes highly negative (-0.35) at 19/ (the highest angle of
attack reached for this oscillation) for the 8/ case. At maximum lift coefficient the moment
is -0.48, in comparison the clean case has a minimum moment coefficient of -0.38 at 22/, a
21% change.

72
2.5 0.1

-0.0

Pitching Moment Coefficient


2.0

-0.1
Lift Coefficient

1.5

-0.2

1.0
-0.3

0.5 2D
-0.4
Clean Re=1 million,
Grit f = 1.8 Hz, 14 ± 10° Sin
0.0 -0.5
-5 0 5 10 15 20 25 30 0 5 10 15 20 25
Angle of Attack, deg Angle of Attack, deg

Figure 51. 2D unsteady lift and pitching moment clean and LEGR, f = 1.8 Hz.

The lift and pitching moment at 1.8 Hz case for both the clean and LEGR applied cases
for 14/ mean angle are presented in figure 51. Both the clean and grit applied data are shown
together for comparison and 14/ mean angle was chosen since in both cases the maximum
lift coefficient occurred in this range of angles of attack. As noted before, the maximum lift
coefficient is higher for the clean case and occurs at a larger angle of attack. The slight
change in the lift curve slope is attributed to the application of the grit, as seen from the
steady state data. The LEGR applied lift also shows a stall at 16/ angle of attack and then
a reattachment of the boundary layer occurs and the lift increases at 22/, this is unstable and
does not occur for all oscillations (an average would still include this phenomena but the
second lift spike would be smaller). The hysteresis loop for the LEGR applied case is larger
and extends into lower angles of attack, meanwhile the clean hysteresis loop closes at 5/
angle of attack. The quarter chord pitching moment coefficient comparison of clean and grit
is shown in the same figure. As can be seen, the grit applied moment is much more negative
at lower angles of attack than the clean case. The moment spikes at 16/, a few degrees before

73
stall occurs, to a value of -0.48 and it is approximately the same value for 5/. The clean case
shows a much higher pitching moment and it does not occur until 22/.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5 2D
Clean Re=1 million,
Grit f = 1.8 Hz, 14 ± 10° Sin
0.0
0 1 2 3
Angle of Attack, deg

Figure 52. 2D lift-time comparison, clean and LEGR, f = 1.8Hz.

To show the detail of the oscilating cycle, the clean and grit cases were plotted with
respect to time of the oscillation. The data were shifted so that the starting angle of attack
was the same, this was necessary since the model moved through few cycles to establish the
flow field and the experimental setup could not start the data acquisition at the same angle
of attack. As can be seen from figure 52, both the clean and the grit applied curves are not
exactly a sinusoidal wave. The closer the lift plot looks like a true sine wave, the smaller
the hysteresis loops become (no difference between the upswing and the downswing). In this
presentation it easy to identify the cycles in which the lift with the grit applied did reach a
lower maximum value than the clean data. The slight shift between the plots indicates that
the grit data stalled earlier in time, therefore it also stalled at a lower angle of attack since
the cams used for the clean and grit applied were the same. The hysteresis lift for the clean

74
data is less repeatable, since in certain cases an earlier boundary layer reattachment can
occur. The grit applied data, however, show a definite minimum lift at the bottom of the
hysteresis loop.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5 2D
Clean Re=1 million,
Grit f = 0.6 Hz, 14 ± 10° Sin
0.0
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 53. 2D lift coefficient comparison, clean and LEGR, f = 0.6Hz.

A comparison between the clean and grit for the low frequency case is shown in
figure 53. With the application of the grit the lift coefficient usually drops significantly,
especially for the low frequency cases, where there is less outside momentum brought into
the boundary layer and the flow behaves more like steady state. The clean case shows a
much higher lift coefficient at approximately 1.9 for the 14/ mean angle, the LEGR applied
only reaches a lift coefficient of 1.5. The LEGR data also stall earlier at 13/ angle of attack,
in comparison to the clean a 18/. The stall angle of attack and the lift coefficient dropped
by approximately 25% in comparison to the clean data. The hysteresis loops look similar in

75
both cases and are almost the same size. The lift coefficient does not return to the same
value as the upswing with the LEGR applied until approximately 5/ angle of attack, while
the clean data reverts back at 8/, a 3/ change.

2.5

2.0
Lift Coefficient

1.5

1.0

Clean
Grit

0.5 2D
Re=1 million,
f = 1.8 Hz, 8 ± 10° Sin
0.0
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 54. 2D lift coefficient comparison, clean and LEGR, f = 1.8 Hz.

76
2.5

2.0
Lift Coefficient

1.5

1.0

Clean
Grit

0.5 2D
Re=1 million,
f = 0.6 Hz, 8 ± 10° Sin
0.0
-5 0 5 10 15 20 25 30
Angle of Attack, deg

Figure 55. 2D lift coefficient comparison, clean and LEGR, f = 0.6 Hz.

In the figures 54 and 55 the comparison of the clean and LEGR is shown for the 2D data
at 8/ mean angle of attack for 1.8 Hz and 0.6 Hz, respectively. The high frequency case,
figure 54, shows that the hysteresis loops for the LEGR applied are almost double in size of
the clean case and extend into lower angles of attack. The maximum lift coefficient reached
by both the clean and grit applied cases is 1.9 but the hysteresis (downswing) value is 1.6 for
the clean and 0.8 for the LEGR. The close up point of the hysteresis loops have moved by
5/, with the grit applied cases reverting to linear curve at 3/ angle of attack. The slope of the
lift curve slope with the grit applied departs earlier from a linear dependence. There is a
noticeable change at 10/ angle of attack for the grit and approximately 15/ for the clean.
The lower frequency changes the shape of the hysteresis loops, as shown in figure 55.
The hysteresis loops for the clean data are now relatively small and occur for a small angle
of attack range from 10/ to 18/. With the application of the LEGR, the size of the loops

77
increases and their extent is longer by 4-5/. The maximum lift coefficient for the clean data
is 1.8 and 1.45 for the grit applied case, a 19% change. The stall angle of attack has dropped
by 2/ from 15/ angle of attack for the clean data. The hysteresis lift for the clean case is 1.45
and for the grit 0.8, a 19% change for clean and 44% for the grit. The hysteresis lift
coefficient for the LEGR is very flat ( at the same value from stall to reattachment), whereas
the clean data show a decrease from a larger value at stall to a lower value at the
reattachment.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5 Clean
Flat VGs
2D, Re=1 million,
f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 56. 2D lift coefficient comparison, clean and VGS, f = 1.8 Hz.

4.2.2 2D Unsteady Data with Vortex Generators

Figure 56 shows the 2D data clean and vortex generators cases for the 1.8 Hz frequency
at 14/ mean angle. As can be seen from the figure, the application of the vortex generators

78
decreases the size of the hysteresis loop significantly. The vortex generators were not
optimized for the 2D data, but with an optimized set of vortex generators the hysteresis loops
may be eliminated. The maximum lift coefficient reached at 22/ angle of attack by the clean
data is 2.3 and by the vortex generators 2.4, a small change. The hysteresis lift though is 1.8
for the vortex generators and only 1.0 for the clean data, this is a significant change. The
downswing lift coefficient is approximately 40% higher than for the clean case.
Whereas the application of the LEGR increases the pressure drag by a few percent close
to stall, the vortex generators do not affect the pressure drag coefficient significantly. The
changes in pressure drag are not large, since the values for the unsteady clean data are already
high. Also the pressure drag is not accurate for the low angles of attack, since it cannot be
resolved with the few taps available at the leading edge. The wake surveys could not be
taken for the unsteady data, since the instrumentation does not have a quick enough response
and also the problem of connecting the right wake to the right angle of attack as the model
oscillates would have to be considered. Only after stall, where the drag values are large, is
the pressure drag somewhat accurate and for those angles of attack the application of grit or
vortex generators does not change the drag much.

79
2.5

2.0
Lift Coefficient

1.5

1.0

Clean
Flat VGs
0.5
2D, Re=1 million,
f = 1.8 Hz, 14 ± 10° Sin
0.0
0 1 2 3
Time, seconds

Figure 57. 2D Lift-time, clean and vortex generators, f = 1.8 Hz.

In order to easily compare the clean and vortex generators data a time plot is shown in
figure 57. As seen from the figure the maximum lift coefficient for both the clean and vortex
generators at 14/ is the same, which suggests that the maximum lift coefficient for the
oscillation cycle occurs at 20/mean angle. The clean curve does not look like a sine wave,
it has a strong peak and a slow return. The vortex generators change the shape of the wave,
so that it begins to resemble a sine wave. When the hysteresis loop is not present the time
plot of the lift coefficient is close to a perfect sine wave, as seen from other experiments
presented later.

80
0.1

Pitching Moment Coefficient -0.0

-0.1

-0.2

-0.3

-0.4 Clean 2D, Re=1 million,


Flat VGs
f = 1.8 Hz, 14 ± 10° Sin

-0.5
0 10 20 30
Angle of Attack, deg

Figure 58. 2D moment coefficient, clean and Flat VGS, f = 1.8 Hz.

The quarter chord pitching moment coefficient at 1.8 Hz for clean data together with the
vortex generators case is shown in figure 58. The pitching moment is most affected at the
stall angle of attack by the application of the vortex generators. Also through the low angle
of attack range on the downswing from approximately 5/ to 12/, the moment with the vortex
generators present is lower by about 0.05 from the clean data. It becomes more negative
between 20/ to 23/ angle of attack and has a minimum value of -0.5 for the vortex generators
and -0.32 for the clean, a 56% change.

81
2.5

2.0
Lift Coefficient

1.5

1.0

0.5 Clean
Flat VGs
2D, Re=1 million,
f = 0.6 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 59. 2D lift coefficient, clean and Flat VGS, f = 0.6 Hz.

In figure 59 the 0.6 Hz frequency case with the applied vortex generators is shown. For
the 14/ mean angle of attack, the vortex generators data obtain a higher lift coefficient of 2.2
than the clean data at 1.9, a 16% increase. The stall angle of attack changes only by 0.5/ and
the overall flow characteristics after stall look similar, even though the clean data obtain
lower lift. The stall is more rounded on the clean data in comparison to the vortex generators
results, where the stall is sharper. The hysteresis loops with the application of the vortex
generators decrease less for the low frequency case than for the higher one. In the 1.8 Hz
case the vortex generators can bring into the boundary layer more of the high momentum
fluid from the outer flow. The hysteresis lift for the clean case is approximately 1.3 and 1.6
for the vortex generators. As mentioned before the low frequency cases after stall always
show slight irrepeatability. The reattachment point for upswing and downswing is the same
for both cases.

82
2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Grit 2D, Re=1 million,


Flat VGs - Grit f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 60. 2D lift coefficient, clean and Flat VGS, f = 1.8 Hz.

The vortex generators were applied to both clean model conditions and with the LEGR
in order to see if it was possible to offset the effects of the grit with the vortex generators.
The effect of the vortex generators was much lower than hoped for on the gritted model.
Figure 60 represents the high frequency case for 14/ mean angle. As can be seen from the
figure, the lift coefficient with the vortex generators and the grit is almost exactly the same.
It changes from 2.15 for the grit to 2.25 for the vortex generators and grit. With grit and
vortex generators the lift coefficient shows a stall which occurs at 22/ angle of attack and
differs from the grit by only 2/. The hysteresis loops are also close in size through the angle
of attack range from 13/ to 22/ with a downswing lift of 0.9. The reattachment point for the
vortex generators data occurs at a slightly higher angle of attack of 4/.

83
2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Grit 2D, Re=1 million,


Flat VGs - Grit f = 1.8 Hz, 14 ± 10° Sin
0.0
0 1 2 3
Time, seconds

Figure 61. 2D lift- time comparison, clean and Flat VGS, f = 1.8 Hz.

The similarities between the flows are visible from figure 61 where the lift coefficient
is plotted with respect to time for both the vortex generators with and without the grit. The
maximum lift coefficient and the time response are almost the same for the two flows. The
application of the vortex generators to the grit do not have an effect on the maximum lift
coefficient for the 14/ mean angle. The small “hiccups” at the low lift values correspond to
a phenomena mostly found in 8/ mean angle of the clean data, where the downswing lift
coefficient becomes higher than the upswing. This pattern is caused by the boundary layer
lagging behind the potential solution for the individual angles of attack. At these frequencies
the flow is assumed to be quasi steady but the boundary layer does not always keep up with
the oscillations. The changes in the boundary layer thickness or the displacement thickness
cause the changes in the lift curves.

84
0.1

Pitching Moment Coefficient -0.0

-0.1

-0.2

-0.3
Grit
Flat VGs - Grit

-0.4 2D, Re=1 million,


f = 1.8 Hz, 14 ± 10° Sin
-0.5
0 10 20 30
Time, seconds

Figure 62. 2D pitching moment coefficient, clean and Flat VGS, f = 1.8 Hz.

The application of the vortex generators to the gritted model decreases the size of the
hysteresis loop in the pitching moment after stall as shown in figure 62. The minimum
moment coefficient changes less than 10% but the extent of the highly negative pitching
moment is enlarged. The range of angle of attack is halved with the application of the vortex
generators. There is a slight lowering of the pitching moment at the low angle of attack from
4/ to 10/.

4.2.3 2D and 3D Unsteady Comparison

In the previous sections the differences between 2D and 3D steady state flow were
explored, also unsteady flow characteristics were looked at in a 2D configuration. From
those results it can be concluded that the steady state and unsteady in 2D vary significantly,

85
and so the 3D unsteady behavior will also change from that of 2D. The 2D lift curve slope
is always much higher than any of the 3D six stations data across the span of the model.
Also as noted before, the stall angle of attack increases for each of the six station in
comparison to the 2D steady state. All these characteristics of the 3D steady flow are found
in the unsteady regime. The 3D oscillating data for clean and grit applied model
configurations is summarized in Appendix B, tables B8 to B31.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5

z=0.075 Clean, Re=1 million,


2D f = 0.6 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 63. Comparison of 2D and 3D lift, clean, f = 0.6 Hz, z=0.075.

Figure 63 shows a comparison of 2D and wall section clean data for the low frequency
case of 0.6 Hz and 14/±10/ oscillations. As can be seen from the figure, the hysteresis loop
is slightly smaller for the 3D case. A lift curve slope of 0.075 is the same through the low
angle of attack range for both 2D and 3D cases. The 2D results exhibit slightly more
curvature in the lift curve at angles of attack higher than 12/. The 3D case does not depart

86
from its linear slope until 18/ angle of attack. The maximum lift coefficient is approximately
1.6 for both cases and occurs at 18/ for the 2D and at 22/ for z=0.075.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5

z=0.075 Clean, Re=1 million,


2D f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 64. Comparison of 2D and 3D lift, clean, f = 1.8 Hz, z=0.075.

For comparison the high frequency case is presented in figure 64. As seen in the low
frequency case the hysteresis loop is smaller in size for the 3D case, but the lift coefficient
is also lower. The 2D lift coefficient is 2.2 and the 3D wall section attains only the value of
2. Even though the wall section is far away from the tip of the airfoil model approximately
29 inches, it is affected by the 3D flow characteristics. The angle of attack of the maximum
lift has changed from 22/ for the 2D to 24/ for the wall station. The hysteresis loop is
smaller as noted before and the lift on the downswing is 1.05 for the 2D and 1.35 for the 3D
wall station. A 52% drop in lift coefficient occurred for the 2D data between the maximum
lift coefficient and the hysteresis one, in comparison only a 32% change occurred for the wall

87
station. The lift curve slope for z=0.075 through the low angle of attack range up to 10/ is
almost equal to 2D case but at higher angles of attack the slope decreases slightly due to the
tip vortex effect.

0.1
Pitching Moment Coefficient

-0.0

-0.1

-0.2

-0.3

-0.4
z=0.075 Clean, Re=1 million,
2D f = 1.8 Hz, 14 ± 10° Sin
-0.5
0 10 20 30
Angle of Attack, deg

Figure 65. Comparison of 2D and 3D clean moment, wall section, f=1.8 Hz, z=0.075.

Figure 65 shows the quarter chord pitching moment coefficient for the 1.8 Hz frequency
case for both the 2D and 3D wall station. The main difference between the 2D and 3D wall
case is that the moment coefficient behaves differently close to stall. In the 3D case the
moment coefficient drops very sharply at approximately 25/ from a constant value of -0.1
and in the downswing section of the curve the moment is on average -0.05. The hysteresis
loops are almost nonexistent for the wall section. In the 2D case the pitching moment
exhibits a hysteresis loop from approximately 20/ to 24/, the minimum value of the moment
is -0.35 equal to that of 3D wall section value. The 2D hysteresis loop can be attributed to

88
an earlier onset of stall. For angles of attack less than 20/ the characteristics of the pitching
moment are very similar for both 2D and wall section and they trace the 2D curves relatively
closely for both the high and low frequencies. The differences in the pitching moment for
the low frequency occur after 15/ angle of attack, where the wall section reaches a higher
pitching moment coefficients by approximately 0.05.

0.8

0.7
Pressure Drag Coefficient

0.6

0.5

0.4

0.3

0.2 z=0.075
2D

0.1 Clean, Re=1 million,


f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 66. Comparison of 2D and 3D pressure drag, clean, f = 1.8 Hz, z=0.075.

The 2D and wall station pressure drag coefficients are shown in figure 66, for the high
frequency. As with the moment coefficient there is little change between the wall section
pressure drag and the 2D data. For the 2D the drag coefficient, like the moment, changes at
approximately 20/ and the pressure drag coefficient increases to 0.6 through a range of 5/.
In the case of 3D wall section data the range narrows to an almost instantaneous increase to
the same maximum value. Also, the pressure drag is higher for the wall section, especially

89
between 10/ and 20/ angle of attack, most likely due to a higher hysteresis lift experienced
by the section.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5

z=0.925 Clean, Re=1 million,


2D f = 0.6 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 67. Comparison of 2D and 3D lift, clean, f = 0.6 Hz, z=0.925.

The z=0.925 station which is only 3 inches from the tip, experiences the highest induced
flow characteristics as can be seen from figure 67, presented here for the low frequency case
of 0.6 Hz. The slope of the lift curve slope is much lower for the tip station than the 2D data,
also the y intercept differs by approximately 0.5. The slope of the 2D case is 0.083 and that
of 3D tip station is 0.060. The tip station lift curve slope is the lowest of all the 6 sections,
the highest slope (close to 2D) is that of the wall section in the previous figures. The
hysteresis loop is very small for z=0.925. The maximum lift coefficient for the tip section
is close to 1.6 which occurs at 22/ angle of attack and the hysteresis lift is approximately 1.4.
In the 2D case the maximum lift coefficient is 1.9 at 18/ and the hysteresis value is 1.3.

90
There is a 32% drop in lift for the 2D and the z=0.925 section experiences only a 12%
decrease. The stall angle of attack increases by 4/ from that of 2D value.
The pitching moment coefficient, not shown, for the 0.6 Hz case is only slightly more
negative than the 2D by 0.05, especially between 10/ and 20/ angle of attack. The 2D and
3D tip station curves at low frequency are together up to 10/ angle of attack and after 20/.
The pitching moment shows almost no hysteresis loops for both the cases. The minimum
moment coefficient for 2D is -0.18 and for the 3D tip station -0.20.

0.8

0.7
Pressure Drag Coefficient

0.6

0.5

0.4

0.3

z=0.925
0.2
2D

0.1 Clean, Re=1 million,


f = 0.6 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 68. Comparison of 2D and 3D pressure drag, clean, f = 1.8 Hz, z = 0.925.

The pressure drag coefficient in figure 68 shows that the tip station sees much higher
pressure drag than the 2D data. Though out the range of angles of attack covered in an
oscillation the pressure drag for the tip section is approximately 2 times higher than the 2D
value at the same angle. At 12/ angle of attack, as an example, the 2D pressure drag

91
coefficient is 0.07 and the 3D tip station has a coefficient of 0.16, a 128% increase occurs.
That increase is due to the effects on the tip vortex of the flow around the wing. The
hysteresis loops are very small in the pressure plot for the tip station, and the 2D has some
“instabilities” showing but no real hysteresis loops.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5

z=0.925 Clean, Re=1 million,


2D f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 69. Comparison of 2D and 3D lift, clean, f = 1.8 Hz, z = 0.925.

The high frequency case for the tip station and the 2D clean model condition are shown
in figure 69. At 1.8 Hz, the hysteresis loop for the 2D clean case is more than 2 times larger
than for the tip station. The maximum lift coefficient for 2D is 2.4 and occurs at 22/ angle
of attack. The corresponding lift coefficient on the downswing was 0.9, that represents a
62% charge in lift. The highest lift obtained by the 3D section is 1.7 at 24/ and the hysteresis
lift is approximately 1.4, that corresponds to only a 17% drop. The lift curve slope for the

92
high frequency case is 0.087 and 0.056, respectively for the 2D and the tip station, the same
as the low frequency case.

0.1
Pitching Moment Coefficient

-0.0

-0.1

-0.2

-0.3

-0.4
z=0.925 Clean, Re=1 million,
2D f = 1.8 Hz, 14 ± 10° Sin
-0.5
0 10 20 30
Angle of Attack, deg

Figure 70. Comparison of 2D and 3D pitching moment, clean, f = 1.8 Hz, z = 0.925.

The pitching moment coefficient is shown for the previous condition in figure 70. On
the average, the values of the coefficient are approximately the same for the 2D and the tip
station. The pitching moment for the 3D section is lower for the angles of attack less than
10/ and has less of a difference between the upswing and downswing values at these angle
of attack. At 10/ angle of attack the 2D pitching moment is -0.02 to -0.11 and the 3D tip
moment is between -0.07 to -0.10. For the angles between 10/ and 20/, the upswing 3D
curve has a lower quarter chord pitching moment coefficient. The minimum coefficient
occurs close to the stall angle of attack and is -0.35 for the 2D and -0.25 for the tip station.

93
0.8

0.7
Pressure Drag Coefficient
0.6

0.5

0.4

0.3

z=0.925
0.2
2D

0.1 Clean, Re=1 million,


f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 71. Comparison of 2D and 3D clean pressure drag, f = 1.8 Hz, z = 0.925.

The pressure drag for 1.8 Hz has a particular kink at the stall angle of attack for the 2D
data, as shown in figure 71. Close to the stall angle the 2D condition has a drag coefficient
that is the same for both upswing and downswing, this phenomena does not repeat for any
other data, it is limited only to this frequency and mean angle. As before the pressure drag
coefficient is larger for the 3D tip station as seen in the 0.6 Hz case, but in the 1.8 Hz case
the increase in not through the whole angle of attack range but only from approximately 10/
to 20/ angle of attack. The maximum pressure drag is the same for 2D and 3D tip station and
has a value of 0.61. For example at 18/ angle of attack the pressure drag coefficient is 0.14
for the upswing and 0.11 for the downswing in the 2D data case and 0.3 for the upswing and
0.2 for the downswing in the 3D tip case. The drag increases between the upswing and
downswing at this angle of attack by 21% and 25% for the 2D and 3D tip station,
respectively.

94
4.2.4 3D Unsteady Data for All Model Conditions

2.5

2.0
Lift Coefficient

1.5

1.0

0.5 Clean
Flat VGs
Flat - Steady z=0.075, Re=1 million,
Clean - Steady f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 72. 3D steady and unsteady comparison, f=1.8 Hz, z = 0.075.

As seen in previous plots the vortex generators increase the maximum lift coefficient.
Also in 2D testing the hysteresis loops with the vortex generators became smaller. The
vortex generators were applied to the 3D model under oscillating conditions for the same
cases as the clean. As can be seen in figure 72, in which the clean wall data for 1.8 Hz case
is shown, the vortex generators have a major effect on the lift coefficient hysteresis. For
comparison, the steady state clean data and the steady state vortex generators data are plotted
together with the clean wall station 14/ mean angle oscillation and the corresponding
unsteady data with the vortex generators. As can be seen from the figure, the clean steady
state data obtain a maximum lift coefficient of 1.5 at approximately 16/ angle of attack. The
maximum lift coefficients for both the steady state and oscillating cases with the vortex

95
generators are the same and equals to 2.1. As noted before the vortex generators increase
the lift and stall angle of attack by approximately 35% in comparison with the clean model
condition. The vortex generators do not increase the lift coefficient in the unsteady testing
for the 1.8 Hz frequency. They do change the behavior of the hysteresis loops. In most
oscillating cases with the vortex generators applied the hysteresis loops are extremely small.
For example at maximum lift for the 14/ mean angle case the clean wall station has a
hysteresis lift of 1.3 and the vortex generators applied case reaches a value of 2.05. The
difference between the maximum lift of 2.1 and hysteresis lift of 2.05 for the vortex
generators applied is nominal. This is a great development, since with the vortex generators
there is no hysteresis loop and the lift is the same no matter of which part of the cycle it falls
on. With no hysteresis loop there is very little change in the loading of the model, also the
penalty of the downswing has been eliminated.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5 Clean
Flat VGs z=0.925, Re=1 million
Flat - Steady
Clean - Steady f = 1.8 Hz, 14 ± 10° Si
0.0
0 10 20 30
Angle of Attack, deg

Figure 73. 3D steady and unsteady comparison, f = 1.8 Hz, z = 0.925.

96
The tip station data is shown in figure 73, for the same conditions as the previous figure.
As before the lift coefficient for the clean steady data is much lower than the oscillating
cases. The maximum lift coefficient for steady state is 1.2 for the clean case and 1.7 for the
vortex generators applied case. The steady state with the vortex generators applied did not
attain the oscillating lift coefficient maximum. It was approximately 11% lower than the
unsteady cases with or without the vortex generators. The hysteresis loops at the tip station
are generally small, and additionally the vortex generators have decreases them even more.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Clean z=0.075, Re=1 million,


Flat VGs f = 0.6 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 74. 3D lift clean and VGs, f = 0.6 Hz, z = 0.075.

The time oscillation comparison for the clean and Flat vortex generators cases is shown
in figure 74. These data are for 0.6 Hz frequency case at the wall station. The maximum
lift coefficient that the vortex generators attain is 2.1 and the clean data have a maximum lift

97
coefficient of 1.8. The maximum lift coefficient in the low frequency case increases by 11%
in comparison to the clean results. The lift curve slope for the vortex generators applied data
is linear through a large angle of attack range, up to 20/, meanwhile the clean data depart
from the linear dependence at approximately 12/ angle of attack. There is a slight change
in the y intercept due to the changes in the boundary layer thickness and displacement
thickness caused by the presence of the vortex generators.

2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Clean z=0.075, Re=1 million,


Flat VGs
f = 0.6 Hz, 14 ± 10° Sin
0.0
0 1 2 3 4 5 6
Time, seconds

Figure 75. A time comparison of 3D clean and VGs, f=0.6 Hz, z = 0.075.

To see the extent of the hysteresis loop more easily, the lift can be plotted against time,
as shown in figure 75. The vortex generators applied lift has almost a perfect sine wave
pattern, but the clean data does not follow the same curve. The clean data are missing the
upper parts of the sine wave, that is where the hysteresis occurs. Also it can be seen that the
three individual oscillations with the vortex generators applied have the same maximum and

98
the same shape making the oscillations repeatable. In comparison, the oscillations for the
clean data do not repeat exactly, since the flow at high angles of attack or after stall is not
stable and its behavior is not repeatable. The slight shift in the low angle of attack is
accidental due to experimental setup.

0.1
Pitching Moment Coefficient

-0.0

-0.1

-0.2

-0.3

-0.4 z=0.075, Re=1 million,


Clean
Flat VGs f = 0.6 Hz, 14 ± 10° Sin
-0.5
0 10 20 30
Angle of Attack, deg

Figure 76. 3D moment clean and VGs, f = 0.6 Hz, z = 0.075.

The pitching moment coefficient for the wall station is shown in figure 76 for the low
frequency case with and without the vortex generators. The moment with the vortex
generators does not show the characteristic spike around the stall angle of attack, which
occurs in the clean data. The value of the moment coefficient is almost constant with the
vortex generators at -0.09. In comparison, the clean wall station data have a slightly higher
moment coefficient on the downswing part of the cycle. The minimum value is
approximately -0.2 at 22/.

99
0.8
Clean
Flat VGs
0.7
z=0.075, Re=1 million,
f = 0.6 Hz, 14 ± 10° Sin
Pressure Drag Coefficient
0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 10 20 30
Angle of Attack, deg

Figure 77. 3D pressure drag clean and VGS, f = 0.6 Hz, z = 0.075.

Additionally the pressure drag coefficient, presented in figure 77 for the z=0.075 case
with and without vortex generators, shows that the increase in the drag is minimal for the
whole range of angles of attack tested. Also the high pressure drag spike at approximately
the stall angle of attack for the clean data disappears with the application of the vortex
generators. For example the pressure drag at 10/ for the clean data is 0.06 on average
between the upswing and downswing and the vortex generators have a coefficient of 0.09.
But at 20/ angle of attack both the clean and vortex generators data have the same drag
coefficient of 0.22.

100
2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Clean z=0.075, Re=1 million,


Flat VGs f = 1.8 Hz, 14, 20 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 78. 3D lift clean and VGs, f = 1.8 Hz, z = 0.075.

The 1.8 Hz frequency case is represented in Figure 78 for both the 14/ and 20/ mean
angles of attack at the wall station with and without the vortex generators. As for the low
frequency with the vortex generators data for 14/ do not show the hysteresis loop and obtain
the same lift coefficient as the clean data. The slope of the lift coefficient curve changes
slightly with the application of the vortex generators to 0.075 from 0.073 for the clean
results. Since the hysteresis loops for the 14/case are small, the lift time plot shows a
sinusoidal curve with the same shape as the forcing function. The 20/ case has a maximum
lift coefficient of 2.3 at 29/, 9% higher that the clean. Sharper stall characteristics are
exhibited by data for the vortex generators case. The hysteresis loop for 20/ mean angle does
exist in both cases, clean and vortex generators applied. The hysteresis lift for the clean data
at 20/ has an average value of 0.9 and the vortex generators case has a value of 1.3, therefore
the hysteresis loop for the vortex generators is approximately 15% smaller.

101
0.1

Pitching Moment Coefficient -0.0

-0.1

-0.2

-0.3

-0.4
Clean z=0.075, Re=1 million,
Flat VGs f = 1.8 Hz, 14 ± 10° Sin
-0.5
0 10 20 30
Angle of Attack, deg

Figure 79. 3D moment clean and VGs, f = 1.8 Hz, z = 0.075.

The vortex generators applied pitching moment for the 14/ mean angle does not have the
sharp drop close to the stall angle of attack, as seen from figure 79. The minimum value at
22/ is approximately -0.1 for the vortex generators and -0.3 for the clean case. The moment,
for example, at 12/ is -0.1 for the vortex generators applied upswing and -0.07 for the
downswing in comparison the clean case has a moment of -0.09 for the upswing and -0.02
for the downswing. The hysteresis loop for the vortex generators case is smaller, as seen
from the above pitching moment coefficients, the loop is approximately half of the size.
The pressure drag with the vortex generators stays almost the same as the clean data at
all frequency, especially at the 14/ mean angle. It is higher by 0.05 from 3/ to 18/ angle of
attack and after 18/the downswing values are lower by 0.05. The pressure drag reaches a
maximum value of 0.33 for the VGS and 0.65 for the clean case; the maximum pressure drag
decreased by 49% for the wall station at 14/ mean angle with the 1.8 Hz frequency.

102
2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Clean z=0.925, Re=1 million,


Flat VGs f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 80. 3D lift clean and VGs, f = 1.8 Hz, z = 0.925.

The tip station as mentioned before showed much smaller hysteresis loops from those of
the wall station due to the induced vortex effects. As can be seen from figure 80, the vortex
generators still decrease the hysteresis loop; the same occurred in the wall station cases. The
flow characteristics around the tip station are more unsteady or turbulent, than for the stations
further from the tip of the model. The effect of the vortex generators is not perfect (the noise
in the data) at the tip station only. The tip station sees the main tip vortex influence and with
it the instabilities which the vortex produces. The individual vortex generators close to the
tip have to compete with the induced vortex, which is being shed from the tip of the model.
That provides more “noise” in the data. The maximum lift coefficient reached by both the
clean and vortex generators applied case is approximately 1.7. For example at 20/ the
upswing lift is 1.4 for both cases and the hysteresis lift is 1.25 for the vortex generators and
1.05 for the clean case, a 10% and 25% change from the upswing for the vortex generators
applied and clean, respectively. The lower frequency of 0.6 Hz also has a smaller hysteresis

103
loop with the vortex generators applied. The vortex generators add some irregularities to the
data, so that it becomes less repeatable. Some of the irregularities are due to the
experimental setup and others to the flow characteristics at the low frequency, since most of
the problems disappear at 1.3 Hz.

0.1

-0.0
Pitching Moment Coefficient

-0.1

-0.2

-0.3

-0.4
Clean z=0.925, Re=1 million,
Flat VGs f = 1.8 Hz, 14 ± 10° Sin
-0.5
0 10 20 30
Angle of Attack, deg

Figure 81. 3D moment clean and VGs, f = 1.8 Hz, z = 0.925.

The quarter chord pitching moment coefficient is shown in figure 81 for the same case
as above. The moment with the vortex generators applied follows the clean upswing plot,
with the minimum coefficient for both of -0.3. The clean moment has a hysteresis loop of
approximately twice the size of the vortex generators one. At 15/ angle of attack, the
pitching moment coefficient is -0.13 for the upswing for both the clean and vortex generators
cases but for the downswing the values are -0.05 for the clean and -0.1 for the vortex
generators results.

104
0.8

0.7
Pressure Drag Coefficient
0.6

0.5

0.4

0.3

0.2 Clean
Flat VGs

0.1 z=0.925, Re=1 million,


f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 82. 3D k/c=0.0019 Cl vs ", ±5.5/, Tred~0.022

The pressure drag coefficient with and without vortex generators for the tip station at 1.8
Hz frequency is shown in figure 82. As with the lift and pitching moment coefficients the
application of the vortex generators reduces the hysteresis loop for the drag coefficient also
at the tip station. The behavior of the pressure drag is very similar to the low frequency case.
The maximum pressure drag coefficient is equal to 0.62 for both the clean and vortex
generators cases. The drag with the vortex generators applied has values almost equal to the
upswing clean data.

105
2.5

2.0
Lift Coefficient

1.5

1.0

0.5
Clean z=0.925, Re=1 million,
Curled VGs f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 83. 3D lift clean and Curled VGs, f = 1.8 Hz, z = 0.925.

The Flat vortex generators have been proven effective in both the steady state and the
unsteady conditions. The Curled vortex generators when tested in the steady state gave
slightly higher lift increase so they were used in unsteady tests. The results are very close
to those of the Flat vortex generators. Since the Flat are easier to make, it would be ideal to
use them. The only difference between the results of these two types is that the Curled
vortex generators have less “noise” in the tip station. The data follow the curves more
smoothly like for the Flat vortex generators at the wall station cases, as can be seen from
figure 83. The maximum lift coefficient is 1.8, the same as for the Flat vortex generators.
Most likely at the tip station a slight curvature to the vortex generators help to keep the flow
going in the right direction, and so gives more consistent results.

106
2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Grit z=0.075, Re=1 million,


Flat VGs - Grit f = 1.8 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 84. 3D lift grit and VGs with grit, f = 1.8 Hz, z = 0.075.

The leading edge grit roughness was also added to the 3D unsteady tests with the vortex
generators for both tip and wall stations. Shown in figure 84 is an example of the wall
station lift for the 1.8 Hz frequency. This severe roughness affects both the clean and vortex
generators flow fields about to the same extent. The roughness only data for this case show
a larger hysteresis loop with a maximum lift coefficient of 2.0 and an angle of attack
reattachment point of 4/. The grit with the vortex generators applied case has a slightly
lower lift coefficient for the whole angle of attack range on the upswing but reaches the same
maximum lift at a slightly higher angle of attack of 22/. The downswing lift reattaches to
the upswing at approximately 7/. The hysteresis loop is smaller with the vortex generators
by about 10%. The lift coefficient for the roughness applied case can drop up to 40%
depending on the conditions from those of clean cases. The quarter chord pitching moment

107
coefficient and the pressure drag coefficient with the vortex generators and roughness
applied are very close to that of only roughness applied presented before.
2.5

2.0
Lift Coefficient

1.5

1.0

0.5

Grit z=0.925, Re=1 million,


Flat VGs - Grit f = 0.6 Hz, 14 ± 10° Sin
0.0
0 10 20 30
Angle of Attack, deg

Figure 85. 3D lift grit and grit with VGS, f = 0.6 Hz, z = 0.925.

The grit and grit with vortex generators applied cases at the tip station see very similar
characteristics to those at the wall station. The vortex generators have more effect on the tip
for 0.6 Hz, as seen from figure 85, than for the higher frequency. The lift coefficient
increased 23% with the application of the vortex generators, and the stall angle of attack has
moved to approximately 22/ instead of 18/ for the roughness only. The hysteresis loop is
also slightly smaller with the vortex generators but the reattachment occurred at
approximately the same angle of attack for both cases between 10-12/. The high frequency
case with vortex generators and grit together, also sees a decrease in the hysteresis loop of
about 15%, but the maximum lift coefficient does not increase from that of the grit only. The

108
pitching moment and pressure drag coefficients change very little with the application of the
vortex generators to the gritted model.

12

10

6
Induced Angle of Attack, deg

-2
14±10° Sin
-4 Re=1 million

-6 z=0.925, Clean
z=0.258, Clean
-8 z=0.925, Flat VGs
z=0.258, Flat VGs

-10
2 4 6 8 10 12 14 16 18 20 22 24

Angle of Attack, deg

Figure 86. 3D induced angle of attack, clean and VGs, f = 1.8 Hz.

4.2.5 Unsteady Integrated Data

The 3D wing characteristics, like the induced angle of attack or induced drag can be
calculated for the unsteady flow. In the case of unsteady flow an average loop has to be
selected for each oscillating condition. The procedure to obtain the induced angle of attack
for 3D unsteady is the same as for steady state. At the same lift the angles of attack for each
section are compared with the 2D unsteady angle, the difference is the induced angle of
attack. Figure 86 represents these comparisons for two spanwise station, in this case of the
tip station (z=0.925) and a station close to the wall (z=0.258). The wall station (z=0.075)

109
was not used because the some wall interference was possible and the small angle changes
could not be resolved. The clean induced angles of attack for the 14/±10/ show that the
induced angle of attack depends highly on the spanwise location. The tip station had larger
induced angles of attack than the wall station. The angles ranged from 4/ to 10/ on the
upswing and down to -5/ on downswing. The negative induced angles of attack correspond
to the negative downwash. The wall station experienced lower induced angles of attack from
2/ to 6/ on the upswing and -2/ on the downswing. Since there is less effect of the tip vortex
at the wall, the negative downwash is much smaller. The application of the vortex generators
did not change the upswing characteristics of the induced angle of attack curves, and had
little effect on the wall section. At the tip section the application of the vortex generators
increased the downswing induced angle to 0/ and the upswing and downswing induced angle
are the same at 18/ angle of attack in comparison to the clean at 12/.

2.0

1.8

1.6

1.4
Integrated Lift Coefficient

1.2

1.0

0.8

0.6 14±10° Sin


Re=1 million
0.4
Clean
Flat VGs
0.2 Steady State

0.0
0 4 8 12 16 20 24

Angle of Attack, deg

Figure 87. 3D integrated lift, f = 1.8 Hz.

110
As for the steady state data the individual lift or drag curves for the six spanwise stations
can be integrated for the 3D unsteady case. The integrated lift coefficient is shown in figure
87 and the integrated induced drag coefficient in figure 88. Both figures contain the clean
and Flat vortex generators data for the 14/ mean angle and the 1.8 Hz frequency. The
integrated lift coefficient for the clean case still exhibits a large hysteresis loop and with the
Flat vortex generators the hysteresis loop decreases. In comparison to the 2D data the
integrated lift coefficients for the 3D are slightly higher and slope is 0.083. The 2D
integrated lift curve slope is 0.108, a 23% change. The lift at the lowest angle of attack of
the cycle is 0.3 for the vortex generators applied and 0.4 for the clean and overall the clean
lift is higher up to 20/. The induced drag when compared to the 2D data is l0% to 20%
higher than the steady state data at high angle of attack. The application of the vortex
generators reduces the drag slightly on the upswing and increases it on the downswing to
approximately 0. Without the vortex generators the induced drag on the downswing gave
negative values, so that it actually created thrust.

111
0.4

14±10° Sin
Re=1 million
0.3 Clean
Flat VGs
Integrated Induced Drag Coefficient

Steady State

0.2

0.1

0.0

-0.1
0 4 8 12 16 20 24

Angle of Attack, deg

Figure 88. 3D integrated drag, f = 1.8 Hz.

112
CHAPTER 5

CONCLUSIONS

The LS(1)-0417MOD airfoil model was tested under an extensive range of conditions
in order to obtain aerodynamic characteristics for 2D and 3D flows with and without vortex
generators and leading edge grit roughness in steady and unsteady conditions. The testing
included angles of attack up to 30/ for steady state tests and ±10/, ±5.5/ pitch oscillations for
0.6, 1.3 ad 1.8 Hz; all these tests were conducted for 0.75, 1.0, 1.25 million Reynolds
numbers.
The 2D tests, which were conducted as a baseline, had a lift coefficient increase of 12%
due to the application of the Flat vortex generators when compared to the clean case,
unfortunately the drag coefficient also increased by 25%. The lift coefficient with the vortex
generators increased because of the “injection” of the higher momentum flow from the outer
layer into the boundary layer. The mixing changes the thickness of the boundary layer and
the displacement thickness, therefore keeping the boundary layer attached at higher angles
of attack. The main penalty of an increase in the lift is the increase in the drag due to the
presence of the vortex generators. The Flat vortex generators became a compromise between
the increase in lift and drag. Even though the Curved vortex generators did increase the lift
more than Flat vortex generators, the drag increase was too high for useful applications,
especially at the low angles of attack.
The flow’s response to the application of the vortex generators, their position or even
their shape was difficult to predict, mostly due to our inability to “calculate” the outer flow
influx into the boundary layer and the response of the boundary layer to that influx. From
the tests performed for the 2D and 3D model conditions, the data show that placing the
vortex generators between 0.07 and 0.15 of the chord increases the lift coefficient

113
significantly while keeping the drag in check. The position of the vortex generators at 0.07
and 0.15 of the chord corresponds to the approximate location of the pressure peak at lower
angles of attack. In placing the vortex generators close to the pressure peak allows the
mixing of the injected outer flow and the boundary layer flow to occur before reaching the
adverse pressure gradient. Additionally for 2D testing only shifting the vortex generators
location rearwards by approximately 10% chord increased the lift coefficient to a higher
value. However, these tests were predominantly for a 3D wing and the vortex generators’
position was optimized for these tests. Therefore the location was kept forwards due to the
lower lift curve slope for the 3D in comparison with the 2D cases. The vortex generators
were successful in increasing the lift coefficient at both the wall station and the tip station
and shifting the stall angle of attack by approximately 5/ to 7/. The Flat vortex generators
increased the lift coefficient at the wall by 26% and at the tip by 33%. The vortex generators
at the tip had to be positioned closer to the leading edge to counteract the effect of the tip
vortex.
The application of the leading edge grit roughness produced a significant loss of lift and
increase in drag, due to the added turbulence in the boundary layer. This type of grit
roughness is very severe and especially at the leading edge of the airfoil it is almost the
height of the boundary layer thickness. The decrease in lift could not be totally countered
with the application of the vortex generators most likely due to their height. It is possible
that taller vortex generators would have more effect in mixing the outer flow, which is less
affected by the roughness, with the boundary layer flow. The Flat vortex generators still
increased the lift coefficient by approximately 25% for both the tip and wall stations and
shifted the stall angle of attack by 5/.
The slope of the 3D integrated data was always lower than the 2D cases, as theory
predicted. The 3D integrated data did not reach the same lift coefficient as the 2D clean case,
but with the application of the vortex generators the 3D integrated case surpassed the 2D case
with vortex generators lift. The application of roughness to the above case caused the
maximum lift coefficient to decrease without vortex generators and increase significantly
over the 2D clean case with the vortex generators applied. The total drag on the wing

114
became equal at high angles of attack no matter what condition was tested. This suggests
that the induced drag from the tip vortex dominates in the calculations and the viscous drag
is only a small part of the total as should be expected from wing theory.
The lift coefficient increased by approximately 50% for the 2D unsteady cases,
unfortunately the unsteady flow experiences large hysteresis loops with hysteresis lift
coefficients 60% lower than the upswing. The increase in the lift is due to the delayed stall
which occurs because of the changes in the boundary layer due to the oscillations. The
boundary layer becomes thinner on the upswing of the oscillation and stays attached until
higher angles of attack. After separation the boundary layer characteristics do not return to
those of the upswing, thus creating the hysteresis loop, in which a different displacement and
boundary layer thickness exist. These changes produce an unsteady stress on the structure
of the wing and have to be incorporated into the design, and they depend on the frequency,
the type of oscillation, the amplitude of the oscillation and the Reynolds number.
The application of the grit roughness increases the size of the hysteresis loops, whereas
the hysteresis lift coefficients are approximately 62% lower than the upswing values. This
sever roughness causes changes in the boundary layer behavior, because the particles produce
the excessive turbulence which in turn increases the thickness of the boundary layer and also
increases its ability to separate. The lift coefficient for the 0.6 Hz frequency case for all
Reynolds numbers shows a much lower maximum lift coefficient with approximately the
same size hysteresis loop. The 1.8 Hz frequency cases have a maximum lift coefficient up
to 100% higher than the steady state case but also have the largest hysteresis experienced,
especially for the ±10/ mean angle of attack. The pitching moment coefficients show a larger
hysteresis loop occurring around the stall angle of attack, which happens earlier for the grit
applied cases.
The application of the vortex generators to the 2D clean case increases the hysteresis lift
slightly for both the frequencies, the 0.6 Hz and 1.8Hz, but also they cause a sharp decrease
in the pitching moment close to the stall angle of attack. The application of the vortex
generators to the gritted model in a 2D setup has very little effect on all the aerodynamic
parameters.

115
The 3D unsteady cases exhibit the same characteristic lower lift curve slope and the
larger stall angle of attack as did the 3D steady state case. In the 3D cases the wall station
in the 0.6 Hz frequency cases obtains almost the same maximum lift coefficient as the 2D
case but at a higher angle of attack. However, this phenomena does not occur in the higher
frequencies cases, where the maximum lift is lower than 2D model condition. Stronger
interaction between the tip vortex and the oscillations occur, especially in the higher
frequency. The pitching moment and pressure drag are not affected much by the 3D
conditions. The tip station exhibits a strong effect of the tip vortex, therefore the lift curve
slope and the maximum lift coefficient are much lower than the wall station and the 2D
cases. From the six stations across the span, only the tip station displays small hysteresis
loops for all the unsteady cases, due to the cross flow from the tip vortex, which is strongest
at the tip station. Even though the tip station exhibits small hysteresis loops, the effect of the
unsteady flow is manifested in the pressure coefficient with a higher maximum peak on the
upswing and lower on the downswing. Also the pressure drag coefficient doubles at the tip
station in comparison to the 2D clean data.
The hysteresis loops were eliminated from the unsteady case with the application of the
vortex generators for all frequency and Reynolds numbers at 8/ and 14/ mean angles of
attack. The pressure distributions for cases with vortex generators applied and clean are the
same for the same values of the lift coefficient, therefore the vortex generators in the 3D
cases are exactly the right height and shape to mix the outer flow with the boundary layer.
Usually only the wall section obtained with the vortex generators in steady state and unsteady
the same lift coefficient, both the lift curve slope and the maximum lift are equal in the two
cases. This suggests that the addition of the correct vortex generators to the unsteady cases,
removes the oscillating behavior of the boundary layer and makes the system behave as
steady state up to stall. After the wall section stall in steady state, the hysteresis loop appears
back in the data as seen from the 20/ mean angle case. The tip station exhibits the same
characteristics, but the vortex generators have a less steady effect, more noise occurs in the
data. The steady state data with the vortex generators at the tip station reached the same stall
angle of attack but a slightly lower maximum lift coefficient. The pitching moment and the

116
pressure drag coefficients are both reduced for both the tip and wall stations with the
application of the vortex generators. It is possible to reduce or remove the hysteresis loop
with the application of the vortex generators from the unsteady data only up to the angle of
attack where the steady state data stalled.
The application of the leading edge grit roughness with the vortex generators returned
the hysteresis loop to almost the same size as the clean 3D cases, especially at the wall
station, however the tip station had still smaller hysteresis loop and higher maximum lift
coefficient than with the grit only.
Overall there was little change in the induced angle of attack due to the application of the
vortex generators to the unsteady flow, however the induced angle of attack was increased
at the tip station for the downswing. The hysteresis loop in the induced angle of attack was
much larger at the wall station than at the tip station, but the induced angles of attack were
generally lower for the wall station. The integrated induced drag was found to be higher for
the steady state than for the unsteady flow, clean or with vortex generators. The vortex
generators cases had the lower induced drag, especially at the higher angle of attack, than the
clean wing condition, which exhibited a large hysteresis loop in the drag.
The data obtained from these experiments provide a large database of pressure
distributions for many model conditions, which before were not available in literature.
These pressure measurements can be used for code verification and development, since many
pressure distributions for different flow characteristics have been obtained. From the six
pressure distributions across the span of the model small nuances in the flow response to
unsteady flow characteristics, applied roughness or vortex generators can be modeled, in
order to obtain more accurate predictions. Also the effect of the tip vortex on the flow across
the span of the model can be modeled with the details found in the pressure distributions.
The six stations across the span provided local lift and pitching moment coefficients which
can be used to model the influence of the tip vortex on the wing as a function of distance
from the tip. These data are important, especially since they showed the influence of the tip
is concentrated in the first two stations closest to the tip on a wing which had a low aspect

117
ratio of 3.3. The inboard stations only display the slightly lower lift curve slope in
comparison to the 2D data.

118
LIST OF REFERENCES

1. Wickens, R.H.: ”Wind Tunnel Investigations of Dynamic Stall of an NACA 0018 Airfoil
Oscillating in Pitch”, Aeronautical Note NAE-AN-27, February 1985.

2. Kim, S.W., Zaman, S.B.M.Q. and Panda, J.: “Numerical Investigation of Unsteady
Transitional Flow over Oscillating Airfoils”, Journal of Fluids Engineering, pp 10-16, March
1995.

3. Olsen, J.: “Unsteady Pressure Measurements on Oscillating Models in European Wind


Tunnels”, AFWAL-TM-80-1-FIBR, March 1980.

4. Levin, V.: “A Synopsis of Boundary Layer Transition”, FSTN 63-1.

5. Burggraf O.R.: “Notes on Unsteady Aerodynamics”, For AAE745, Spring 1987.

6. McCroskey, W.J.: “Inviscid Flowfield of an Unsteady Airfoil”, AIAA Journal, Vol 11,
Num 8, pp 1130-1136, August 1973.

7. Chambers, F.W.: “Synthetically Generated Turbulent Boundary Layer Development and


Structure”, AIAA Journal, Vol 24, Num 12, Dec 1986, pp 1987-1993.

8. Jones, W.P.: “Research on Unsteady Flow”, Journal of Aerospace Sciences, Vol 29, Num
3, March 1962, pp 249-263.

9. Houwink, R., van der Vooren, J.: “Results of an Improved Version of LTRAN2 for
Computing Unsteady Airloads on Airfoils Oscillating in Transonic Flow”, Presented at the
12th Fluid and Plasma Dynamics Conference, July 23-25, 1979, Williamsburg, VA.

10. Reissner, E.: “On the Application of Mathieu Function in the Theory of Subsonic
Compressible Flow Past Oscillating Airfoils”, NACA TN-2363, May 1951.

11. Pope, A.: “On Airfoil Theory and Experiment”, Journal of Aeronautical Sciences,
Vol 15, Num 7, July 1948, pp 407-410.

12. Postel, E.E., Leppert, E.L., Jr.: “Theoretical Pressure Distributions for a Thin Airfoil
Oscillating in Incompressible Flow”, Journal of Aeronautical Sciences, Vol 15, Num 8,
August 1948, pp 486-492.

119
13. Hahneman, E., Freedman, J.C. and Finston, M.: “Stability of Boundary Layers and of
Flow in the Entrance Section of a Channel”, Journal of Aeronautical Sciences, Vol 15,
Num 8, August 1948, pp493-496.

14. Shimizu, Y., Ismaili, E., Kamada, Y., Maeda, T.: “Power Augmentation of a HAWT by
Mie-type Tip Vanes, Considering Wind Tunnel Flow Visualization, Blade-Aspect Ratios and
Reynolds Number”, Wind Engineering, Vol 27, Num 3, 2003, pp 183-194.

15. Lai, J.C.S., Platzer, M.F.: “”Jet Characteristics of a Plunging Airfoil”, AIAA Journal,
Vol 37, Num 12, December 1999, pp 1529-1537.

16. Lee, J.D., Gregorek, G.M.: “Notes on Vortex Generators”

17. Mateescu, D., Abdo, M.: “Unsteady Aerodynamic Solutions for Oscillating Airfoils”,
AIAA 2003-227, Presented at the 41st Aerospace Sciences Meeting and Exhibit, 6-9 January,
2003, Reno, NV.

18. Sellers III, W.L., Jones, G.S., Moore, M.D.: “Flow Control Research at NASA Langley
in Support of High-Lift Augmentation”, AIAA 2002-6006, Presented at Biennial
International Powered Lift Conference, 5-7 November, 2002.

19. Rae, A.J., Galpin, S.A., Fulker, J.: “Investigation into Scale Effects on the Performance
of Sub Boundary Layer Vortex Generators on Civil Aircraft High-Lift Devices”, AIAA 2002-
3274.

20. Patel, M.P., Carver, R., et-al.: “Detection and Control of Flow Separation Using Pressure
Sensors and Micro-Vortex Generators”, AIAA 2002-268, Presented at the 40th Aerospace
Science Meeting and Exhibit, 14-17 January, 2002, Reno, NV.

21. Ashill, P.R., Fulker, J.L. and Hackett, K.C.: “Research at DERA on Sub Boundary Layer
Vortex Generators (SBVGs)”, AIAA 2001-0887, Presented at 39th Aerospace Sciences
Meeting and Exhibit, 8-11 January, 2001, Reno, NV.

22. Ashill, P.R., Fulker, J.L. and Hackett, K.C.: ”Studies of Flows Induced by Sub Boundary
Layer Vortex Generators (SBVGs)”, AIAA 2002-0968, Presented at the 40th Aerospace
Sciences Meeting and Exhibit, 14-17 January, 2002, Reno, NV.

23. Lee, J.S., Kim, C., and Rho, O.H.: “The Modification of Airfoil Shape for Optimal
Aerodynamic Performace on Flapping Airfoils in Low Reynolds Number Flow”,
AIAA 2002-0421, Presented at the 41st Aerospace Sciences Meeting and Exhibit, 6-9
January, 2002, Reno, NV.

120
24. Sunada, S., Kawachi, K., et-al.: “Unsteady Forces on the Two-Dimensional Wing in
Plunging and Pitching Motions”, AIAA Journal, Vol 39, Num 7, July 2001, pp 1230-1239.

25. Klausmeyer, S.M., Papadakis, M., Lin, J.: “A Flow Physics Study of Vortex Generators
on a Multi-element Airfoil”, AIAA 1996-0548, Presented at the 34th Aerospace Sciences
Meeting and Exhibit, 15-18 January, 1996, Reno, NV.

26. Reuster, J.G., Baeder, J.D.: “ Effects of Plunging on Flow Control for Pitching Airfoils”,
AIAA 2001-2976, Presented at the 31st AIAA Fluid Dynamics Conference, 11-14 June, 2001,
Anahein, CA.

27. Wendt, B.J., Reichert, B.A.: “The Modeling of Symmetric Airfoil Vortex Generators”,
AIAA 1996-0807, Presented at the 34th Aerospace Sciences Meeting and Exhibit, 15-18
January, 1996, Reno, NV.

28. Pascazio, M., Autric, J.M., et-al.: ”Unsteady Boundary Layer Measurement on
Oscillating Airfoils - Transition and Separation Phenomena in Pitching Motion”,
AIAA 1996-0035, Presented at the 34th Aerospace Sciences Meeting and Exhibit, 15-18
January, 1996, Reno, NV.

29. Ericsson, L.E.: “Dynamic Stall of Pitching Airfoils and Slender Wings – Similarities and
Differences”, AIAA 1998-0414.

30. Amitay, M., Smith, D.R., et-al.: “Aerodynamic Flow Control over an Unconventional
Airfoil Using Synthetic Jet Actualors”, AIAA Journal, Vol 39, Num 3, March 2001,
pp 361-370.

31. Hoerner, S., Borst, H.: Fluid Dynamic Lift, Published by Mrs. L.A. Horner, 1975.

32. Timmer, W.A., van Rooji, R.P.J.O.M.: “Summary of the Delft University Wind Turbine
Dedicated Airfoils”, AIAA 2003-0352, Presented at the 41st Aerospace Sciences Meeting and
Exhibit, 6-9 January, 2003, Reno, NV.

33. Van Rooji, R.P.J.O.M., Timmer, W.A.: “Roughness Sensitivity Considerations for Thick
Rotor Blade Airfoils”, AIAA 2003-0350, Presented at the 41st Aerospace Sciences Meeting
and Exhibit, 6-9 January, 2003, Reno, NV.

34. Bertagnolio, F., Sørensen, N.N., et-al.: “Conclusions from Comparisons of Numerous
Two Dimensional Airfoil Computations with Experiment”, AIAA 2002-0034, Presented at
the 40th Aerospace Sciences Meeting and Exhibit, 14-17 January, 2002, Reno, NV.
35. Gaunaa, M., Sørensen, N.N.: “Experimental Investigation of Airfoils Subject to
Harmonic Translatory Motions”, AIAA 2002-0035, Presented at the 40th Aerospace Sciences
Meeting and Exhibit, 14-17 January, 2002, Reno, NV.

121
36. Leishman, J.G.: “Challenges in Modeling the Unsteady Aerodynamics of Wind
Turbines”, AIAA 2002-0037, Presented at the 40th Aerospace Sciences Meeting and Exhibit,
14-17 January, 2002, Reno, NV.

37. Fuglsang, P., Bak, C.: “Design and Verification of the New Risø-A1 Airfoil Family for
Wind Turbines”, AIAA 2001-0028, Presented at the 39th Aerospace Sciences Meeting and
Exhibit, 11-14 January, 2001, Reno, NV.

38. Jones, R.T.: “The Unsteady Lift on a Wing of Finite Aspect Ratio”, NACA 681, 1940.

39. Björck, A.: “Dynamic Stall and Three-Dimensional Effects”, FFA-TN-1995-31, Sweden.

40. Rueger, M.L.: “An Experimental Investigation of the Effect of Vortex Generators an
Aerodynamic Characteristics of a NACA 0021 Airfoil Undergoing Large Amplitude Pitch
Oscillations:, OSU Thesis, 1988.

41. Kusunose, K., Neng, J.Y.: “Vortex Generator Installation Drag on an Airplane Near Its
Cruise Condition”, AIAA 2003-0932.

42. Kiedaisch, J.W., Acharya, M.: “Investigation of Unsteady Separation over Pitching
Airfoils at High Reynolds number”, AIAA 1998-2975.

43. Lin, J.C.: “Control of Turbulent Boundary Layer Separation using Micro-Vortex
Generators”, AIAA 1999-3404.

44. Mishtawy, J.E.: “An Investigation into the Unsteady Lift Characteristics of Four Airfoils
of Unique Shape”, OSU Thesis, 2003.

45. Pope, A., Harper, J.J.: Low Speed Wind Tunnel Testing, John Wiley and Sons, Inc., New
York, 1966.

46. Schlichting, H.: Boundary Layer Theory, McGraw-Hill Inc., New York, 1979.

47. Dommasch, D., Sherby S., and et-al.: Airplane Aerodynamics, Pitman Publishing Corp.,
New York, 1961.

48. Smetana, F., Summey, D., et-al.: “Light Aircraft Lift, Drag and Moment Prediction - a
Review and Analysis”, North Carolina State University, NASA CR-2523.

49. Von Mises, R.: Theory of flight, McGraw-Hill Inc., New York, 1945.

50. Gormont, R.E.: “A mathematical Model of Unsteady Aerodynamics and Radial Flow
for Application to Helicopter Rotors”, USAAMRDL, Technical Report 72-67, 1973.

122
Appendix A: Measured Model Coordinates

Tap Number Chord Station Ordinate


1 1.00124 0.00199
2 0.97451 0.00067
3 0.94938 0.00192
4 0.89975 0.00055
5 0.84941 -0.00486
6 0.79992 -0.01239
7 0.69985 -0.03084
8 0.60001 -0.04793
9 0.49968 -0.05882
10 0.39949 -0.06319
11 0.29962 -0.06336
12 0.19980 -0.05859
13 0.14952 -0.05360
14 0.09994 -0.04645
15 0.07528 -0.04179
16 0.05002 -0.03530
17 0.03508 -0.03034
18 0.02027 -0.02411
19 0.00854 -0.01718
20 0.00000 0.00636
21 0.00778 0.02679
22 0.02007 0.04181
23 0.03459 0.05423

Continued

Table A1. LS(1)-0417MOD 3D Surface Pressure Taps, Non-Dimensional Coordinates for


z=0.258, z=0.592, and z=0.925

123
Table A1 continued

24 0.04963 0.06392
25 0.07482 0.07667
26 0.09948 0.08606
27 0.14971 0.09851
28 0.19971 0.10514
29 0.29967 0.10709
30 0.39932 0.10208
31 0.49943 0.09508
32 0.59973 0.08504
33 0.69964 0.06959
34 0.79972 0.05016
35 0.84949 0.03938
36 0.89948 0.02905
37 0.94954 0.01783
38 0.97423 0.01167

124
Tap Number Chord Station Ordinate
1 1.00161 0.00027
2 0.97474 0.00050
3 0.95005 0.00166
4 0.89964 0.00023
5 0.84983 -0.05160
6 0.79951 -0.01294
7 0.69972 -0.03137
8 0.59954 -0.04838
9 0.49968 -0.05897
10 0.39989 -0.06329
11 0.29972 -0.06360
12 0.19971 -0.05877
13 0.14998 -0.05377
14 0.09993 -0.04652
15 0.07498 -0.04172
16 0.05009 -0.03550
17 0.03517 -0.03058
18 0.02021 -0.02438
19 0.00860 -0.01731
20 0.00000 0.00639
21 0.00766 0.02673
22 0.01965 0.04142
23 0.03475 0.05415
24 0.04958 0.06383
25 0.07492 0.07673
26 0.09999 0.08601
27 0.15004 0.09850
28 0.19984 0.10507

Continued

Table A2. LS(1)-0417MOD 3D Surface Pressure Taps, Non-Dimensional Coordinates for


z=0.075, z=0.408, and z=0.742

125
Table A2 Continued

29 0.29969 0.10673
30 0.39964 0.10184
31 0.49966 0.09507
32 0.59972 0.08511
33 0.69960 0.06942
34 0.79926 0.05011
35 0.85016 0.03949
36 0.89957 0.02865
37 0.95004 0.01752
38 0.97458 0.01153

126
Appendix B: Summary of Data

Grit Pattern Re x 10-6 Clmax Cdmin Cmo dCl/d"


Clean 0.75 1.62 @ 16.4/ 0.0080 -0.0850 0.112
k/c=0.0019 0.75 1.14 @ 10.3/ 0.0140 -0.0682 0.109
Clean 1.00 1.58 @ 14.4/ 0.0086 -0.0826 0.113
Flat 1.00 1.77 @ 14.1/ 0.0104 0.117
k/c=0.0019 1.00 1.19 @ 10.0/ 0.0138 -0.0708 0.111
Flat + LEGR 1.00 1.53 @ 13.1/ 0.0156 0.118
Clean 1.25 1.61 @ 14.4/ 0.0080 -0.0804 0.111
k/c=0.0019 1.25 1.22 @ 10.0/ 0.0120 -0.0692 0.111

Table B1. LS(1)-0417MOD 3D Steady State Parameters Summary for 2D

Tred Re x 10-6 f Clmax "max Cl dec Cm inc Cm dec


0.031 0.75 0.56 1.83 18.8 1.59 -0.1027 -0.0850
0.071 0.75 1.29 1.95 20.4 1.22 -0.1488 -0.1064
0.099 0.74 1.82 2.24 23.1 1.33 -0.2688 -0.1180
0.023 1.01 0.56 1.86 19.1 1.30 -0.1291 -0.1478
0.054 1.01 1.34 1.94 20.7 1.21 -0.1666 -0.0953
0.074 1.00 1.82 1.99 19.8 1.16 -0.1429 -0.0978
0.019 1.24 0.57 1.83 18.3 1.40 -0.1258 -0.1100
0.043 1.24 1.32 1.90 20.6 1.16 -0.1698 -0.1018
0.058 1.24 1.82 1.95 20.9 1.24 -0.2013 -0.1107

Table B2. LS(1)-0417MOD 3D, 2D Unsteady, Clean, ±5.5/

127
Tred Re x 10-6 f Clmax "max Cl dec Cm inc Cm dec
0.031 0.74 0.55 1.42 13.5 0.86 -0.1239 -0.0806
0.075 0.74 1.34 1.61 15.8 0.84 -0.2483 -0.0973
0.101 0.74 1.82 1.84 17.0 0.89 -0.3345 -0.1393
0.023 1.00 0.56 1.31 12.0 0.84 -0.0653 -0.0836
0.055 1.00 1.34 1.54 14.7 0.78 -0.2351 -0.0774
0.077 0.99 1.85 1.63 15.7 0.72 -0.2511 -0.0854
0.019 1.24 0.56 1.32 12.1 0.79 -0.0723 -0.0791
0.045 1.24 1.37 1.46 13.8 0.74 -0.1473 -0.0707
0.059 1.25 1.79 1.55 14.9 0.74 -0.1719 -0.0804

Table B3. LS(1)-0417MOD 3D, 2D Unsteady, LEGR, ±5.5/

Tred Re x 10-6 f Clmax "max Cl dec Cm inc Cm dec


0.031 0.75 0.56 1.90 17.7 1.04 -0.0931 -0.0718
0.074 0.74 1.32 2.29 21.9 1.12 -0.2343 -0.1016
0.103 0.74 1.82 2.78 24.9 1.27 -0.4673 -0.1282
0.023 0.99 0.56 1.86 17.7 1.09 -0.0978 -0.0951
0.058 1.00 1.37 2.09 20.6 1.06 -0.1748 -0.0996
0.078 0.99 1.85 2.52 23.4 1.13 -0.3904 -0.1347
0.018 1.24 0.57 1.83 17.2 1.05 -0.0872 -0.0848
0.045 1.24 1.34 2.03 20.4 1.00 -0.2206 -0.1181
0.062 1.24 1.85 2.29 22.4 0.99 -0.3452 -0.1224

Table B4. LS(1)-0417MOD 3D, 2D Unsteady, Clean, ±10/

128
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.76 0.56 1.58 15.9 0.85 -0.2673 -0.0895
0.074 0.74 1.32 2.09 18.6 0.83 -0.4150 -0.1006
0.104 0.73 1.82 2.49 20.7 1.09 -0.4414 -0.1917
0.024 0.98 0.56 1.51 14.8 0.92 -0.1940 -0.0979
0.058 0.98 1.37 1.88 17.0 0.84 -0.3146 -0.0873
0.078 0.98 1.85 2.25 22.9 1.01 -0.4326 -0.1665
0.019 1.22 0.57 1.40 13.1 0.83 -0.0858 -0.0835
0.045 1.22 1.32 1.81 16.0 0.81 -0.2852 -0.0825
0.061 1.22 1.79 2.02 17.6 0.76 -0.3066 -0.0851

Table B5. LS(1)-0417MOD 3D, 2D Unsteady, LEGR, ±10/

129
Z Grit Pattern Re x 10-6 Clmax Cmo dCl/dC"
0.075 Clean 0.75 1.58 @ 19.5/ -0.0721 0.079
0.075 k/c=0.0019 0.75 1.18 @ 14.2/ -0.0660 0.079
0.075 Clean 1.00 1.52 @ 17.4/ -0.0696 0.080
0.075 k/c=0.0019 1.00 1.16 @ 13.2/ -0.0641 0.078
0.075 Clean 1.25 1.51 @ 19.3/ -0.0741 0.078
0.075 k/c=0.0019 1.25 1.19 @ 14.2/ -0.0682 0.078
0.258 Clean 0.75 1.54 @ 19.5/ -0.0773 0.079
0.258 k/c=0.0019 0.75 1.14 @ 14.2/ -0.0653 0.075
0.258 Clean 1.00 1.48 @ 17.4/ -0.0787 0.078
0.258 k/c=0.0019 1.00 1.11 @ 13.2/ -0.0677 0.075
0.258 Clean 1.25 1.54 @ 18.2/ -0.0768 0.076
0.258 k/c=0.0019 1.25 1.17 @ 14.2/ -0.0692 0.075
0.408 Clean 0.75 1.50 @ 18.3/ -0.0825 0.074
0.408 k/c=0.0019 0.75 1.14 @ 14.3/ -0.0705 0.074
0.408 Clean 1.00 1.51 @ 18.4/ -0.0802 0.074
0.408 k/c=0.0019 1.00 1.10 @ 13.3/ -0.0674 0.073
0.408 Clean 1.25 1.54 @ 18.3/ -0.0766 0.076
0.408 k/c=0.0019 1.25 1.09 @ 13.1/ -0.0719 0.072
0.592 Clean 0.75 1.44 @ 18.3/ -0.0838 0.071
0.592 k/c=0.0019 0.75 1.11 @ 14.3/ -0.0709 0.068
0.592 Clean 1.00 1.46 @ 18.4/ -0.0786 0.070
0.592 k/c=0.0019 1.00 1.06 @ 14.2/ -0.0661 0.068
0.592 Clean 1.25 1.49 @ 18.3/ -0.0765 0.070
0.592 k/c=0.0019 1.25 1.05 @ 13.1/ -0.0708 0.067
0.742 Clean 0.75 1.38 @ 19.3/ -0.0822 0.067
0.742 k/c=0.0019 0.75 1.02 @ 14.5/ -0.0683 0.063
0.742 Clean 1.00 1.35 @ 18.4/ -0.0769 0.064
0.742 k/c=0.0019 1.00 0.98 @ 13.2/ -0.0637 0.063
0.742 Clean 1.25 1.38 @ 18.1/ -0.0717 0.066
0.742 k/c=0.0019 1.25 1.03 @ 14.3/ -0.0654 0.064
0.925 Clean 0.75 1.23 @ 19.2/ -0.0583 0.049
0.925 k/c=0.0019 0.75 0.87 @ 14.4/ -0.0544 0.050
0.925 Clean 1.00 1.20 @ 18.3/ -0.0613 0.053
0.925 k/c=0.0019 1.00 0.82 @ 13.2/ -0.0504 0.052
0.925 Clean 1.25 1.22 @ 18.0/ -0.0558 0.054
0.925 k/c=0.0019 1.25 0.88 @ 14.3/ -0.0530 0.051

Table B6. LS(1)-0417MOD 3D Steady State Parameters Summary for 3D

130
Z Grit Pattern Re x 10-6 Clmax Cmo dCl/dC"
0.075 Clean 1.00 2.05 @ 24.5/ 0.083
0.075 k/c=0.0019 1.00 1.48 @ 17.2/ 0.078
0.258 Clean 1.00 2.03 @ 24.5/ 0.080
0.258 k/c=0.0019 1.00 1.47 @ 17.2/ 0.078
0.408 Clean 1.00 1.78 @ 23.3/ 0.076
0.408 k/c=0.0019 1.00 1.49 @ 18.2/ 0.076
0.592 Clean 1.00 1.75 @ 23.3/ 0.073
0.592 k/c=0.0019 1.00 1.46 @ 18.2/ 0.073
0.742 Clean 1.00 1.54 @ 21.2/ 0.064
0.742 k/c=0.0019 1.00 1.34 @ 18.1/ 0.064
0.925 Clean 1.00 1.47 @ 25.2/ 0.051
0.925 k/c=0.0019 1.00 1.16 @ 18.0/ 0.057

Table B7. LS(1)-0417MOD 3D Steady State Parameters Summary for 3D with VGs

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.030 0.75 0.55 1.88 23.2 1.28 -0.1460 -0.1096
0.073 0.74 1.34 2.20 27.4 1.18 -0.2782 -0.1582
0.097 0.74 1.79 2.36 27.8 1.09 -0.2542 -0.1696
0.022 0.99 0.56 1.92 21.9 1.20 -0.1795 -0.0784
0.055 0.99 1.34 2.13 25.7 0.96 -0.2172 -0.0946
0.076 0.99 1.88 2.26 26.8 0.82 -0.2503 -0.1516
0.018 1.24 0.56 1.93 21.5 1.22 -0.3046 -0.1131
0.044 1.24 1.34 2.15 23.2 1.55 -0.3360 -0.1959
0.058 1.24 1.82 2.27 26.3 1.11 -0.2935 -0.1436

Table B8. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.075

131
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.032 0.74 0.56 1.71 18.8 0.84 -0.2089 -0.1127
0.076 0.75 1.34 2.17 22.4 0.81 -0.3519 -0.1106
0.101 0.75 1.79 2.35 22.5 1.02 -0.4886 -0.1618
0.023 0.99 0.56 1.62 17.9 0.76 -0.2243 -0.0837
0.056 0.99 1.32 1.85 19.8 0.80 -0.2047 -0.0890
0.077 0.99 1.82 2.19 23.1 0.73 -0.3103 -0.1030
0.019 1.25 0.56 1.58 17.7 0.78 -0.2174 -0.0888
0.045 1.25 1.34 1.82 19.4 0.79 -0.2952 -0.0998
0.062 1.24 1.85 2.00 21.3 0.81 -0.2385 -0.1159

Table B9. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.075

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.030 0.75 0.55 1.98 24.8 0.92 -0.2833 -0.1667
0.073 0.74 1.34 2.07 26.7 1.03 -0.3203 -0.1150
0.097 0.74 1.79 2.37 27.8 0.88 -0.2974 -0.1618
0.022 0.99 0.56 1.78 22.8 1.08 -0.1424 -0.1029
0.055 0.99 1.34 1.98 24.5 1.19 -0.1493 -0.1259
0.076 0.99 1.88 2.20 26.8 0.85 -0.2803 -0.1667
0.018 1.24 0.56 1.78 21.8 1.03 -0.1335 -0.0952
0.044 1.24 1.34 1.97 24.5 1.08 -0.2307 -0.1204
0.058 1.24 1.82 2.10 25.5 1.21 -0.1853 -0.1314

Table B10. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.258

132
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.032 0.74 0.56 1.62 18.7 0.70 -0.2244 -0.0810
0.076 0.75 1.34 1.80 21.3 0.73 -0.2706 -0.1107
0.101 0.75 1.79 1.98 24.7 0.85 -0.2939 -0.1524
0.023 0.99 0.56 1.45 17.8 0.73 -0.1850 -0.0985
0.056 0.99 1.32 1.68 20.0 0.74 -0.2421 -0.0961
0.077 0.99 1.82 1.87 21.9 0.67 -0.2760 -0.0869
0.019 1.25 0.56 1.41 17.3 0.72 -0.1726 -0.0987
0.045 1.25 1.34 1.58 18.7 0.75 -0.1698 -0.0954
0.062 1.24 1.85 1.79 21.3 0.79 -0.2664 -0.1230

Table B11. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.258

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.031 0.74 0.56 1.90 23.1 0.87 -0.1651 -0.1276
0.075 0.74 1.34 2.17 26.3 0.97 -0.2829 -0.1336
0.101 0.74 1.82 2.31 29.0 1.31 -0.3313 -0.3106
0.023 0.99 0.56 1.82 23.2 1.16 -0.2207 -0.1529
0.056 0.98 1.34 2.12 25.3 0.96 -0.2711 -0.1315
0.075 0.99 1.82 2.22 27.1 1.15 -0.2695 -0.1699
0.018 1.23 0.56 1.79 22.5 1.02 -0.1779 -0.1260
0.044 1.24 1.34 1.94 24.6 1.09 -0.2266 -0.1438
0.059 1.24 1.79 2.06 26.9 0.94 -0.2780 -0.1367

Table B12. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.408

133
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.76 0.56 1.71 19.5 0.69 -0.3250 -0.0749
0.072 0.76 1.32 1.96 22.7 0.79 -0.4237 -0.1174
0.102 0.76 1.88 2.34 26.7 0.89 -0.4420 -0.1727
0.023 1.00 0.56 1.55 18.8 0.73 -0.2598 -0.0855
0.056 1.00 1.34 1.95 21.4 0.75 -0.3875 -0.1206
0.074 1.00 1.79 2.18 24.5 0.82 -0.4914 -0.1305
0.019 1.23 0.56 1.55 18.0 0.78 -0.2628 -0.0963
0.045 1.23 1.34 1.78 20.6 0.77 -0.3750 -0.1049
0.061 1.23 1.82 2.05 22.5 0.75 -0.3981 -0.1039

Table B13. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.408

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.031 0.74 0.56 2.06 24.4 1.39 -0.2073 -0.1280
0.075 0.74 1.34 2.41 28.1 1.21 -0.4266 -0.1440
0.101 0.74 1.82 2.57 29.5 2.29 -0.2972 -0.5562
0.023 0.99 0.56 2.03 24.2 1.22 -0.3965 -0.1171
0.056 0.98 1.34 2.30 26.8 1.20 -0.4077 -0.1360
0.075 0.99 1.82 2.39 28.5 1.21 -0.3329 -0.1712
0.018 1.23 0.56 1.97 23.2 1.29 -0.2938 -0.1404
0.044 1.24 1.34 2.20 25.5 1.19 -0.4295 -0.1213
0.059 1.24 1.79 2.31 27.4 1.26 -0.3736 -0.1526

Table B14. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.592

134
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.76 0.56 1.74 19.5 0.78 -0.2150 -0.0929
0.072 0.76 1.32 2.10 23.2 1.04 -0.3842 -0.1902
0.102 0.76 1.88 2.44 27.7 1.15 -0.4542 -0.2530
0.023 1.00 0.56 1.59 18.8 0.77 -0.2126 -0.0991
0.056 1.00 1.34 1.93 21.4 0.78 -0.2489 -0.1214
0.074 1.00 1.79 2.24 24.5 0.87 -0.3497 -0.1382
0.019 1.23 0.56 1.54 18.8 0.81 -0.2240 -0.1028
0.045 1.23 1.34 1.83 20.6 0.81 -0.2779 -0.1198
0.061 1.23 1.82 2.10 23.3 0.76 -0.3564 -0.1132

Table B15. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.592

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.030 0.74 0.56 1.85 24.9 1.35 -0.1518 -0.1278
0.073 0.74 1.34 2.17 27.5 1.22 -0.2179 -0.1282
0.099 0.74 1.82 2.37 29.6 1.84 -0.2432 -0.3121
0.022 1.00 0.55 1.79 23.3 1.25 -0.1188 -0.0935
0.054 1.01 1.34 1.99 26.5 1.17 -0.2053 -0.1195
0.071 1.01 1.79 2.16 28.0 1.24 -0.1858 -0.1684
0.018 1.25 0.56 1.70 23.4 1.26 -0.1336 -0.1092
0.042 1.25 1.32 1.95 25.1 1.21 -0.1557 -0.1195
0.058 1.25 1.82 1.09 27.0 1.23 -0.1683 -0.1363

Table B16. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.742

135
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.75 0.57 1.58 19.7 0.87 -0.2043 -0.1281
0.072 0.75 1.34 1.85 23.8 0.79 -0.2241 -0.1229
0.098 0.75 1.82 2.20 26.4 1.09 -0.2596 -0.2141
0.023 1.00 0.56 1.49 18.2 0.85 -0.1574 -0.1087
0.054 1.00 1.32 1.74 21.7 0.85 -0.1973 -0.1275
0.074 0.99 1.82 1.97 23.7 0.85 -0.2159 -0.1287
0.018 1.26 0.56 1.39 17.7 0.83 -0.1563 -0.0957
0.042 1.26 1.32 1.63 19.9 0.77 -0.1669 -0.1037
0.058 1.26 1.82 1.81 22.8 0.82 -0.2078 -0.1310

Table B17. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.742

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.030 0.74 0.56 1.72 24.3 1.12 -0.2612 -0.1584
0.073 0.74 1.34 2.04 27.5 1.14 -0.3499 -0.1293
0.099 0.74 1.82 2.19 29.6 1.66 -0.3857 -0.2446
0.022 1.00 0.55 1.65 22.5 1.30 -0.2609 -0.1522
0.054 1.01 1.34 1.83 26.5 1.08 -0.3090 -0.1051
0.071 1.01 1.79 2.04 28.0 1.17 -0.3677 -0.1505
0.018 1.25 0.56 1.62 23.4 1.24 -0.2752 -0.1762
0.042 1.25 1.32 1.79 25.1 1.15 -0.2840 -0.1265
0.058 1.25 1.80 1.98 28.1 1.22 -0.3738 -0.1515

Table B18. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.925

136
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.75 0.57 1.33 18.9 0.76 -0.1985 -0.0937
0.072 0.75 1.34 1.81 23.8 0.68 -0.3435 -0.0965
0.098 0.75 1.82 2.04 26.4 0.93 -0.3908 -0.1822
0.023 1.00 0.56 1.25 19.2 0.76 -0.1985 -0.0891
0.054 1.00 1.32 1.56 21.8 0.78 -0.2639 -0.1143
0.074 0.99 1.82 1.80 24.5 0.77 -0.2853 -0.1231
0.018 1.26 0.56 1.19 17.7 0.77 -0.1669 -0.0984
0.042 1.26 1.32 1.46 20.5 0.72 -0.2326 -0.0802
0.058 1.26 1.82 1.64 22.8 0.70 -0.2774 -0.1011

Table B19. LS(1)-0417MOD 3D, Unsteady, LEGR, ±10/, z=0.925

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.031 0.75 0.56 1.80 21.5 1.39 -0.1003 -0.1515
0.074 0.75 1.32 1.96 24.1 1.16 -0.2090 -0.1149
0.102 0.75 1.82 2.08 25.3 1.29 -0.2089 -0.1547
0.024 1.00 0.56 1.82 20.9 1.30 -0.0961 -0.1116
0.055 1.00 1.32 1.88 22.8 1.15 -0.1359 -0.0932
0.074 1.00 1.79 2.05 24.0 1.34 -0.1673 -0.1487
0.019 1.24 0.56 1.86 21.7 1.37 -0.1516 -0.1267
0.044 1.24 1.32 1.91 22.7 1.15 -0.1480 -0.1119
0.061 1.25 1.82 2.03 23.2 1.21 -0.1692 -0.1025

Table B20. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.075

137
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.74 0.56 1.50 17.1 0.88 -0.1589 -0.1085
0.073 0.75 1.32 1.63 18.6 0.94 -0.2068 -0.1325
0.101 0.75 1.82 1.69 19.3 1.09 -0.1900 -0.2293
0.024 1.00 0.56 1.46 16.8 0.86 -0.1997 -0.1044
0.056 1.00 1.34 1.55 17.9 0.83 -0.1692 -0.0981
0.076 1.00 1.82 1.68 18.8 1.08 -0.1895 -0.1743
0.018 1.25 0.55 1.42 16.1 0.99 -0.1249 -0.1308
0.044 1.25 1.34 1.59 17.6 0.84 -0.1895 -0.1482
0.060 1.25 1.82 1.60 18.4 0.81 -0.1779 -0.1153

Table B21. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.075

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.031 0.75 0.56 1.78 21.9 1.20 -0.1309 -0.1113
0.074 0.75 1.32 1.91 22.7 1.16 -0.1343 -0.0985
0.102 0.75 1.82 2.01 24.9 1.46 -0.2354 -0.1923
0.024 1.00 0.56 1.75 20.9 1.15 -0.0722 -0.1411
0.055 1.00 1.32 1.80 22.8 1.01 -0.1268 -0.1003
0.074 1.00 1.79 1.93 24.0 1.17 -0.2163 -0.1667
0.019 1.24 0.56 1.77 21.1 1.03 -0.1382 -0.1047
0.044 1.24 1.32 1.82 21.9 1.03 -0.1228 -0.1074
0.061 1.25 1.82 1.92 23.2 1.06 -0.2100 -0.1408

Table B22. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.258

138
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.74 0.56 1.40 17.1 0.79 -0.1752 -0.1070
0.073 0.75 1.32 1.49 18.6 0.89 -0.2277 -0.1115
0.101 0.75 1.82 1.62 19.23 0.99 -0.1730 -0.2087
0.024 1.00 0.56 1.33 15.7 0.84 -0.0882 -0.0996
0.056 1.00 1.34 1.47 17.7 0.83 -0.1369 -0.1071
0.076 1.00 1.82 1.55 17.8 0.87 -0.1103 -0.1239
0.018 1.25 0.55 1.36 16.1 0.92 -0.1433 -0.1165
0.044 1.25 1.34 1.43 16.8 0.77 -0.1057 -0.0930
0.060 1.25 1.82 1.52 17.9 0.75 -0.1385 -0.0924

Table B23. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.258

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.034 0.78 0.56 1.84 22.1 1.15 -0.1804 -0.0864
0.083 0.78 1.34 1.87 23.2 1.14 -0.1327 -0.0930
0.113 0.78 1.82 2.00 25.2 1.47 -0.2753 -0.2155
0.027 1.01 0.57 1.80 21.6 1.20 -0.1454 -0.0875
0.063 1.02 1.34 1.96 23.6 1.13 -0.3550 -0.1055
0.086 1.02 1.82 2.04 2438 1.33 -0.4416 -0.1448
0.021 1.26 0.56 1.76 21.0 1.24 -0.1687 -0.1192
0.050 1.26 1.32 1.94 22.6 1.10 -0.3741 -0.1082
0.069 1.26 1.82 1.95 24.0 1.08 -0.3492 -0.1354

Table B24. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.408

139
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.034 0.77 0.55 1.40 16.4 0.78 -0.1333 -0.0780
0.084 0.77 1.34 1.63 18.8 0.81 -0.3126 -0.1236
0.117 0.77 1.88 1.80 19.4 1.32 -0.3726 -0.3124
0.027 1.01 0.57 1.40 16.3 0.75 -0.1929 -0.0846
0.062 1.02 1.32 1.55 17.4 0.77 -0.2185 -0.1084
0.087 1.02 1.85 1.74 19.0 1.03 -0.3547 -0.1589
0.022 1.26 0.56 1.33 16.4 0.75 -0.1728 -0.0976
0.049 1.26 1.29 1.47 17.3 0.71 -0.2721 -0.0866
0.067 1.26 1.79 1.65 18.3 0.79 -0.2332 -0.1133

Table B25. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.408

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.031 0.74 0.56 1.81 22.1 1.28 -0.1182 -0.0900
0.075 0.74 1.34 2.05 23.8 1.31 -0.2132 -0.1161
0.101 0.74 1.82 2.24 25.2 1.57 -0.3125 -0.1563
0.023 0.99 0.56 1.81 22.1 1.27 -0.2089 -0.1165
0.056 0.98 1.34 2.00 23.6 1.21 -0.2203 -0.1002
0.075 0.99 1.82 2.13 24.6 1.42 -0.2856 -0.1446
0.018 1.23 0.56 1.74 21.0 1.28 -0.1235 -0.0849
0.044 1.24 1.34 1.97 23.1 1.31 -0.2890 -0.1269
0.059 1.24 1.79 2.10 24.0 1.26 -0.2562 -0.1406

Table B26. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.592

140
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.031 0.76 0.56 1.43 17.3 1.34 -0.1981 -0.1125
0.072 0.76 1.32 1.68 18.8 0.87 -0.2059 -0.1264
0.102 0.76 1.88 1.73 19.4 1.27 -0.2345 -0.1815
0.023 1.00 0.56 1.40 16.8 0.79 -0.1608 -0.0929
0.056 1.00 1.34 1.58 17.9 0.83 -0.1866 -0.1229
0.074 1.00 1.79 1.73 19.0 1.04 -0.2245 -0.1765
0.019 1.23 0.56 1.36 15.1 0.89 -0.1090 -0.1205
0.045 1.23 1.34 1.48 18.0 0.85 -0.1862 -0.1247
0.061 1.23 1.82 1.59 18.4 0.81 -0.1616 -0.1141

Table B27. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.592

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.032 0.76 0.55 1.75 22.3 1.41 -0.1090 -0.1330
0.076 0.76 1.29 1.91 23.9 1.35 -0.1422 -0.1037
0.111 0.76 1.88 2.01 25.0 1.69 -0.1374 -0.1503
0.026 1.00 0.57 1.69 21.8 1.28 -0.1253 -0.0624
0.061 1.01 1.37 1.90 23.4 1.39 -0.2126 -0.1328
0.083 1.01 1.88 1.89 24.4 1.57 -0.1361 -0.2177
0.019 1.27 0.55 1.67 21.2 1.40 -0.1071 -0.1743
0.048 1.28 1.37 1.81 22.8 1.46 -0.1386 -0.1529
0.062 1.27 1.79 1.89 24.0 1.35 -0.1910 -0.1360

Table B28. LS(1)-0417MOD 3D, Unsteady, Clean, ±5.5/, z=0.742

141
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.033 0.76 0.56 1.35 18.4 0.90 -0.1382 -0.1091
0.079 0.76 1.32 1.52 19.0 1.04 -0.1838 -0.1483
0.110 0.76 1.85 1.60 22.5 0.88 -0.2810 -0.1250
0.025 1.01 0.56 1.33 17.4 0.96 -0.1307 -0.1244
0.058 1.01 1.32 1.48 18.6 1.06 -0.1772 -0.1497
0.080 1.01 1.79 1.54 18.9 1.05 -0.1505 -0.1197
0.020 1.25 0.55 1.28 17.0 0.97 -0.1075 -0.1118
0.047 1.25 1.32 1.46 17.9 0.91 -0.1384 -0.1050
0.065 1.25 1.82 1.53 18.9 1.15 -0.1393 -0.1942

Table B29. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.742

Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec


0.032 0.76 0.55 1.66 22.3 1.27 -0.2646 -0.1760
0.076 0.76 1.29 1.83 24.3 1.27 -0.3074 -0.1591
0.111 0.76 1.88 1.86 25.0 1.62 -0.2904 -0.2541
0.026 1.00 0.57 1.61 22.1 1.20 -0.2652 -0.1202
0.061 1.01 1.37 1.74 23.4 1.29 -0.3147 -0.1708
0.083 1.01 1.88 1.83 24.4 1.42 -0.3192 -0.2220
0.019 1.27 0.55 1.56 21.8 1.31 -0.2439 -0.1797
0.048 1.28 1.37 1.69 22.9 1.37 -0.2736 -0.1974
0.062 1.27 1.79 1.78 24.0 1.18 -0.3046 -0.1183

Table B30. LS(1)-0417MOD 3D, Unsteady, Clean, ±10/, z=0.925

142
Tred Re x 10-6 T Clmax "max Cl dec Cm inc Cm dec
0.033 0.76 0.56 1.22 18.4 0.77 -0.1741 -0.0716
0.079 0.76 1.32 1.32 18.5 0.92 -0.2039 -0.1219
0.110 0.76 1.85 1.45 22.5 0.78 -0.2503 -0.1072
0.025 1.01 0.56 1.22 17.4 0.86 -0.1943 -0.1105
0.058 1.01 1.32 1.29 18.4 0.80 -0.1878 -0.0842
0.080 1.01 1.79 1.38 18.9 0.91 -0.2091 -0.0814
0.020 1.25 0.55 1.12 16.4 0.74 -0.1533 -0.0723
0.047 1.25 1.32 1.26 17.9 0.83 -0.1925 -0.0961
0.065 1.25 1.82 1.38 18.9 1.00 -0.2254 -0.1235

Table B31. LS(1)-0417MOD 3D, Unsteady, LEGR, ±5.5/, z=0.925

143
Appendix C: Permissions for reprinting

For Reference 2

Subject: Re: Permission to reprint a figure


To: janiszewska.1@osu.edu
Cc: Philip DiVietro <DiVietroP.Mailman.NYC@asme.org>,
Meghna Patel <MegPatel@asme.org>
Dear Ms Janiszewska,

It is our pleasure to grant you permission to use Figure 4 from "Numerical Investigation of Unsteady
Transitional Flow over Oscillating airfoil" by S.W. Kim , Journal of Fluids Engineering, March 1995,
Vol 117, pp10-16, cited in your letter for inclusion in a PhD. thesis in aeronautical engineering at the
Ohio State University . As is customary, we request that you ensure proper acknowledgment of the
exact sources of this material, the authors, and ASME as original publisher.

In accordance with ASME policy, this permission is contingent upon payment of a royalty fee of
US$20 for 1 figure ($20.00 for the first figure/table, $10 thereafter). This is solely charged to non-
authors of the requested ASME papers. We accept payments on all major credit cards such as: Visa,
MasterCard, American Express, Discover, and Diners Club, or by check payable to ASME. Please
send payment to the attention Meghna Patel, ASME Accounting, 22 Law Drive, Fairfield, NJ 07007,
and indicate A/C# 104-78-0000-4303. Should you have any questions regarding payment form or
transfer, please contact Ms. Patel; P: 973-244-2268, F: 973-882-4924; E: megpatel@asme.org.

Thank you for your interest in ASME publications.

Sincerely,

Beth Darchi

Copyrights & Permissions

V: 212-591-7700

F: 212-591-7292

E: darchib@asme.org

Credit card payment ending in 7057 was used.

144
For Reference 31

Subj:Fluid-Dynamic Lift
Date:3/18/04 5:02:15 PM Pacific Standa:.;
From:HMy
To:jmj@aarls.eng.ohio-state.edu
Dear Jolanta M. Janiszewska:

We herewith give permission for the use of the following figures of the book
"Fluid-Dynamic Lift" as long as you
give full recognition in your thesis to the origin of these figures.
Chapter III figure 1 and
chapter XVII figure 4.
Please mail a copy of this page to our address P.O. Box 21992 in Bakersfield CA 93390.
We wish you good luck.

Sincerely,
Hoerner Fluid Dynamics
Mrs. i A Hoerner
LAH/ka

145
For Reference 44

Jolanta,

I have no problem with you using my figures. If possible, could you send me
a .pdf version of your dissertation when you finish? I'd like to see what
you did. I wish you the best of luck on your exam. Keep in touch.

Jason Mishtawy

146
For Reference 36

To: Jolanta Janiszewska <jmj@aarls.eng.ohio-state.edu>


Subject: Re: Permission for reprinting

I would like to use one of the figures from your paper AIAA-2002-0037 (Fig 17) in my
dissertation which I am writing in the Aeronautical and Astronautical Engineering at
OSU.
Jolanta:

I am happy to give you permission. If you would like an electronic copy of the figure,
please let me know and what digital format you'd like it.

Gordon
J. Gordon Leishman, PhD, DSc
Professor of Aerospace Engineering, University of Maryland at College Park,
College Park, Maryland 20742, USA
Phone:(301)405-1126 FAX:(301)314-9001
Email: leishman@eng.umd.edu

147

Vous aimerez peut-être aussi