APPLICABILITY OF SPECIFIC SPEED AND ZWEIFEL COEFFICIENT RECOMMENDATIONS TO LOW HEAD AXIAL HYDRAULIC TURBINES By Zachary Hoyle A THESIS Submitted to Michigan State University In partial fulfillment of the requirements For the degree of Mechanical Engineering – Master of Science 2017 ABSTRACT APPLICABILITY OF SPECIFIC SPEED AND ZWEIFEL COEFFICIENT RECOMMENDATIONS TO LOW HEAD AXIAL HYDRAULIC TURBINES By Zachary Hoyle This thesis work investigates the applicability of literature recommendations regarding specific speed and Zweifel coefficient for low head axial hydraulic turbines with constant blade thickness circular arc blade profiles, and no inlet guide vane. Highest efficiency was observed in Computational Fluid Dynamics (CFD) investigations for the designs running at speeds 12-33% that of the speed recommended by the Cordier line. Designs that followed the Zweifel coefficient recommendations of the literature in range of 0.8-1.1 demonstrated highest efficiency in CFD investigations. To test the applicability of the Cordier line recommendations for machines of this type, designs at differing specifications were tested at six different rotational speeds. Peak efficiency was measured at specific speeds between 1.4 and 1.9, compared to the Cordierrecommended value of 5.5 to 6.1. Next designs with differing values of Zweifel coefficient were simulated at the rotational speed at which highest efficiency was measured for each set of specifications. Zweifel coefficient was altered by changing the axial blade length. It was found that highest efficiency was measured at Zweifel coefficients between 0.8 and 1.1. The designs with higher axial blade lengths had lower Zweifel coefficient, and experienced greater friction losses. Decreasing blade length and increased Zweifel coefficients were correlated with more severe velocity gradients at the leading edge, increased flow deviation and areas of low pressure where cavitation could occur. Designs with the shortest blades (Zwefiel coefficient above 1.2) experienced greater velocity gradients and turbulence at the trailing edge and decreased efficiency. ii ACKNOWLEDGEMENTS I greatly appreciate the support of those who have helped me through the Mechanical Engineering Master’s program at Michigan State University. First, I thank my family for the love and support they have always provided, my father Randy, my mother Jill, my brother Randall, and my sister Jacqui. I’d like to thank my committee, Dr. Xinran Xiao, Dr. Abraham Engeda, and Dr. Norbert Mueller for their assistance, patience, and help to complete this work. A special thank you to Dr. Norbert Mueller, who for the past 5 years has provided enormously valuable advice and opportunities which have helped me gain experience in engineering and turbomachinery. All of the experience which I now have on my resume is the result of opportunities which Dr. Mueller has provided. I would like to thank my friends and colleagues who have provided very valuable guidance which helped me complete this thesis, in particular Blake Gower and Thomas Qualman. Thank you to Tim Hinds and Andrew Mccolm for providing the financial support which has allowed me to complete the program. iii TABLE OF CONTENTS LIST OF TABLES ......................................................................................................................... vi LIST OF FIGURES ...................................................................................................................... vii KEY TO SYMBOLS ..................................................................................................................... xi CHAPTER 1. INTRODUCTION AND BACKGROUND ............................................................ 1 1.1. Introduction to Hydroelectric Power .................................................................................... 1 1.2. Hydroelectric Power Challenges .......................................................................................... 2 1.3. Hydroelectric Dam Retrofits ................................................................................................ 3 1.4. Selection of Type of Water Turbine ..................................................................................... 6 1.5. Woven Wheel Hydro Turbine .............................................................................................. 7 1.5.1. Lightweight, Modular Design........................................................................................ 9 1.5.2. Low-cost Additive Manufacturing ................................................................................ 9 1.5.3. Rapid Customization and Manufacturing .................................................................... 10 1.5.4. Outer Shroud Eliminates Tip Leakage and Reduces Fish Mortality ........................... 11 1.6. Definition of Turbine Type Investigated ............................................................................ 12 1.7. Conservation of Energy...................................................................................................... 13 1.8. Definitions of Geometry and Design Parameters ............................................................... 17 1.9. Turbine Cascade Forces ..................................................................................................... 25 1.10. Dimensionless Coefficients .............................................................................................. 28 1.10.1. Flow and Capacity Coefficients ................................................................................ 28 1.10.2. Blade Loading and Head Coefficients ....................................................................... 28 1.10.3. Specific Speed ........................................................................................................... 29 1.10.4. Specific Diameter ...................................................................................................... 30 1.11. Cordier Diagram and Line................................................................................................ 30 1.11.1. Effect of Flow, Loading Coefficients on Cordier Diagram Position ......................... 33 1.11.2. Using Cordier Diagram as a Design Tool ................................................................. 35 1.11.3. Operating below the Cordier Line ............................................................................. 36 1.12. Literature Recommendations on Blade Loading and Flow Coefficients ......................... 37 1.13. Cost Advantages of Designs With High Flow and High Loading Coefficients .............. 39 1.14. Zweifel Lift Coefficient ................................................................................................... 40 1.14.1. Recommended Values of Zweifel Coefficient .......................................................... 45 1.14.2. Industry Trend - Higher Zweifel Coefficients ........................................................... 45 1.14.3. Zweifel Analysis ........................................................................................................ 47 1.15. Cavitation ......................................................................................................................... 49 1.15.1. Cavitation for Composites ......................................................................................... 49 1.16. Discussion of Loss Mechanisms ...................................................................................... 50 1.17. Goals of This Thesis and Description of Work ................................................................ 52 CHAPTER 2. ANALYSIS DESCRIPTION AND METHODS................................................... 55 2.1. Description of Setup ........................................................................................................... 55 2.2. Analysis Description, Description of Specifications.......................................................... 56 2.3. Estimation of Friction Losses ............................................................................................. 57 iv 2.4. CFD Methodology and Assumptions ................................................................................. 64 2.4.1. CFD Overview ............................................................................................................. 64 2.4.2. CFD Methodology and Assumptions .......................................................................... 67 2.5. Description and Inputs for Speed Study............................................................................. 69 2.6. Description and Inputs for Blade Length Study ................................................................. 76 2.7. Mesh Independence Study ................................................................................................. 80 CHAPTER 3. CFD RESULTS AND DISCUSSION ................................................................... 84 3.1. Speed Study CFD Results .................................................................................................. 84 3.2. Discussion of Cordier Recommendations and Location on Diagram ................................ 86 3.3. Discussion of Speed Study Results .................................................................................... 89 3.3.1. Overall Summary......................................................................................................... 89 3.3.2. Flow Visualization and Vectors .................................................................................. 91 3.3.3 Trailing Edge Effects .................................................................................................... 95 3.3.4. Blade Loading.............................................................................................................. 97 3.3.5. Hub and Shroud Effects ............................................................................................. 103 3.3.6. Work Extraction ........................................................................................................ 107 3.3.7. Cavitation................................................................................................................... 110 3.4. Second Study: Blade Length Study.................................................................................. 113 3.5. Blade Length Study Results Discussion........................................................................... 115 3.5.1. Overall Summary....................................................................................................... 115 3.5.2. Flow Visualization and Vectors ................................................................................ 118 3.5.3. Trailing Edge Effects ................................................................................................. 123 3.5.4. Blade Loading............................................................................................................ 126 3.5.5. Hub and Shroud Effects ............................................................................................. 131 3.5.6. Work Extraction ........................................................................................................ 134 3.5.7. Cavitation .................................................................................................................. 139 CHAPTER 4. CONCLUSIONS ................................................................................................. 142 4.1. Speed Study and Cordier Line ......................................................................................... 142 4.2. Blade Length Study and Zweifel Coefficient ................................................................... 143 4.3. Next Steps of the Project .................................................................................................. 144 REFERENCES ........................................................................................................................... 147 v LIST OF TABLES Table 2.1. Case 1 Geometric Parameters and Flow Angles .......................................................... 72 Table 2.2. Case 1 Speed Study Inputs........................................................................................... 74 Table 2.3. Case 2 Speed Study Inputs........................................................................................... 75 Table 2.4. Case 3 Speed Study Inputs........................................................................................... 75 Table 2.5. Case 4 Speed Study Inputs........................................................................................... 75 Table 2.6. Case 5 Speed Study Inputs........................................................................................... 75 Table 2.7. Case 6 Speed Study Inputs........................................................................................... 76 Table 2.8. Blade Length Study Runs and Inputs .......................................................................... 80 Table 3.1. Case 1 Speed Study CFD Results ................................................................................ 84 Table 3.2. Case 2 Speed Study CFD Results ................................................................................ 84 Table 3.3. Case 3 Speed Study CFD Results ................................................................................ 85 Table 3.4. Case 4 Speed Study CFD Results ................................................................................ 85 Table 3.5. Case 5 Speed Study CFD Results ................................................................................ 85 Table 3.6. Case 6 Speed Study CFD Results ................................................................................ 86 Table 3.7. Performance of Cordier Recommendations of Cases in Speed Study ......................... 88 Table 3.8. Case 1 Blade Length Study Results ........................................................................... 113 Table 3.9. Case 2 Blade Length Study Results ........................................................................... 113 Table 3.10. Case 3 Blade Length Study Results ......................................................................... 114 Table 3.11. Case 4 Blade Length Study Results ......................................................................... 114 Table 3.12. Case 5 Blade Length Study Results ......................................................................... 115 Table 3.13. Case 6 Blade Length Study Results ......................................................................... 115 vi LIST OF FIGURES Figure 1.1. Hydroelectric Dam Schematic ...................................................................................... 2 Figure 1.2. Bellaire Dam, Antrim County, Michigan ..................................................................... 5 Figure 1.3. Spillway at the Bellaire Dam, Antrim County, Michigan ............................................ 5 Figure 1.4. Hydroelectric dam retrofit schematic ........................................................................... 6 Figure 1.5. (a) 3-D model of Woven Wheel winding scheme (b) Woven Wheel in mandrel (c) Woven wheel removed from mandrel............................................................................................. 8 Figure 1.6. Woven Wheel tidal turbine being tested inside the tow tank ....................................... 8 Figure 1.7. Modular Woven Wheel compressor unit design .......................................................... 9 Figure 1.8. (a) 3-D printed mandrel. (b) carbon fiber winding. (c) mandrel post-winding after being dipped into epoxy/resin and cured. (d) Woven Wheel after removal from mandrel. ......... 10 Figure 1.9. Turbomachine control volume ................................................................................... 14 Figure 1.10. Example velocity triangle ......................................................................................... 17 Figure 1.11. Axial turbine geometry parameters .......................................................................... 19 Figure 1.12. Turbine blade cascade with velocity triangles.......................................................... 22 Figure 1.13. Pressure distribution across a turbine stage .............................................................. 25 Figure 1.14. Example turbine blade cascade with blade forces .................................................... 26 Figure 1.15. Example efficiency curve for turbine for specific speed 𝑁𝑠 [15] ............................ 30 Figure 1.16. 𝐷𝑠 − 𝛺𝑠 diagram with Cordier line and typical machine type ............................... 32 Figure 1.17. 𝐷𝑠 − 𝛺𝑠 diagram with Voith Hydro turbines with Cordier line ............................. 33 Figure 1.18. 𝐷𝑠 − 𝛺𝑠 diagram with Cordier line, lines of constant flow coefficient 𝜙 ............... 34 Figure 1.19 𝐷𝑠 − 𝛺𝑠 diagram with Cordier line, lines of constant blade loading coefficient 𝜓 .. 34 Figure 1.20. Smith loading vs. Flow coefficient diagram [22] ..................................................... 37 Figure 1.21. Blade loading and flow coefficients for different types of hydraulic turbines [15] . 39 Figure 1.22. Profile loss coefficient vs. s/b ratio resulting from separation and friction [14] ...... 41 Figure 1.23. Idealized pressure distribution with actual pressure distribution for Zweifel coefficient calculation ................................................................................................................... 43 Figure 1.24. Blade length vs. Zweifel coefficient ......................................................................... 47 Figure 1.25. Number of blades vs. Zweifel coefficient ................................................................ 48 Figure 1.26. Tip speed vs. Zweifel coefficient at differing blade lengths .................................... 48 Figure 1.27. Trailing edge of blade with separated boundary layer [32] ...................................... 50 Figure 2.1. Hydroelectric dam retrofit schematic with height positions....................................... 55 Figure 2.2. Design 1j5 generated in BladeGen ............................................................................. 67 Figure 2.3. Example turbine geometry with entry and exit annuli ............................................... 67 Figure 2.4. Process flow chart for Speed Study ............................................................................ 71 Figure 2.5. Velocity triangles for leading and trailing edges of design 1j of the Speed Study (50% span) .............................................................................................................................................. 73 Figure 2.6. Velocity triangles for leading and trailing edges of design 1CORD of the Speed Study (50% span) .......................................................................................................................... 74 Figure 2.7. Blade Length Study process flow chart ...................................................................... 77 Figure 2.8. Velocity triangles for leading and trailing edges of design 2l of the Blade Length Study (50% span) .......................................................................................................................... 78 Figure 2.9. Velocity triangles for leading and trailing edges of design 2l8 of the Blade Length Study (50% span) .......................................................................................................................... 79 vii Figure 2.10. Mesh cell count vs. Power for case 1 mesh Study ................................................... 81 Figure 2.11. Mesh cell count vs. Power for case 2 mesh Study ................................................... 81 Figure 2.12. Mesh cell count vs. Power for case 3 mesh Study ................................................... 82 Figure 2.13. Mesh cell count vs. Power for case 4 mesh Study ................................................... 82 Figure 2.14. Mesh cell count vs. Power for case 5 mesh Study ................................................... 82 Figure 2.15. Mesh cell count vs. Power for case 6 mesh Study ................................................... 83 Figure 3.1 Specific speed 𝛺𝑠 Vs. Efficiency of Speed Study designs ......................................... 87 Figure 3.2. Location of peak efficiency points for each case on 𝛺𝑠vs 𝐷𝑠 diagram with Cordier line................................................................................................................................................. 88 Figure 3.3 Flow coefficient vs. Efficiency of Speed Study designs ............................................. 90 Figure 3.4. Blade loading coefficient vs. Efficiency of Speed Study designs .............................. 90 Figure 3.5. Relative velocity vectors for case 1: Span 50% ......................................................... 91 Figure 3.6. Relative velocity vectors for case 2: Span 50% ......................................................... 91 Figure 3.7. Relative velocity vectors for case 3: Span 50% ......................................................... 92 Figure 3.8. Relative velocity vectors for case 4: Span 50% ......................................................... 92 Figure 3.9. Relative velocity vectors for case 5: Span 50% ......................................................... 92 Figure 3.10. Relative velocity vectors for case 6: Span 50% ....................................................... 93 Figure 3.11. Relative velocity vectors at leading edge for case 2: 50% span ............................... 93 Figure 3.12. Relative velocity vectors at leading edge for case 1: Span 50% .............................. 94 Figure 3.13. Relative velocity vectors at leading edge for case 3: Span 50% .............................. 94 Figure 3.14. Relative velocity vectors at trailing edge for case 2: 50% span ............................... 94 Figure 3.15. Relative velocity vectors at trailing edge for case 1: 50% span ............................... 95 Figure 3.16. Relative velocity vectors at trailing edge for case 3: Span 50% .............................. 95 Figure 3.17. Turbulence kinetic energy contours for case 3: Span 50% ...................................... 96 Figure 3.18. Turbulence kinetic energy contours for case 1: Span 50% ...................................... 96 Figure 3.19. Relative velocity of case 1 Speed Study designs near the trailing edge: Span 50% 97 Figure 3.20. Blade loading diagrams for Speed Study case 2: Span 50% .................................... 99 Figure 3.21. Blade loading diagrams for Speed Study case 1: Span 50% .................................... 99 Figure 3.22. Static pressure contours for Speed Study case 2: Span 50% .................................. 100 Figure 3.23. Static pressure contours for Speed Study case 1: Span 50% .................................. 100 Figure 3.24. Static pressure contours for Speed Study case 3: Span 50% .................................. 101 Figure 3.25. Velocity triangles with relative velocity contours for design 1CORD of the Speed Study: (Span 50%) ...................................................................................................................... 102 Figure 3.26. Velocity triangles with relative velocity contours for design 1j of the Speed Study: (Span 50%).................................................................................................................................. 103 Figure 3.27. Relative velocity of Case 2 Speed Study designs at the trailing edge along span.. 104 Figure 3.28. Relative velocity of Case 1 Speed Study designs at the trailing edge along span.. 104 Figure 3.29. Relative velocity of highest efficiency Speed Study designs at the leading edge along span ................................................................................................................................... 105 Figure 3.30. Relative velocity of highest efficiency Speed Study designs at the trailing edge along span ................................................................................................................................... 106 Figure 3.31. Turbulence kinetic energy of highest efficiency Speed Study designs at the trailing edge along span ........................................................................................................................... 107 Figure 3.32. Absolute circumferential flow velocity for designs in case 2 of Speed Study at the trailing edge along span .............................................................................................................. 108 viii Figure 3.33. Absolute Circumferential flow velocity for designs in Case 3 of Speed Study at the trailing edge along span .............................................................................................................. 108 Figure 3.34. Average absolute circumferential flow velocity from inlet to outlet for designs case 1 of Speed Study ......................................................................................................................... 109 Figure 3.35. Absolute circumferential flow velocity for highest efficiency designs in Speed Study at the trailing edge along span .................................................................................................... 110 Figure 3.36. Areas of low static pressure for case 1 of Speed Study: 50% Span ....................... 111 Figure 3.37. Areas of low static pressure for case 2 of Speed Study: 50% Span ....................... 111 Figure 3.38. Areas of low static pressure for case 3 of Speed Study: 50% Span ....................... 111 Figure 3.39. Areas of low static pressure for case 4 of Speed Study: 50% Span ....................... 112 Figure 3.40. Areas of low static pressure for case 5 of Speed Study: 50% Span ....................... 112 Figure 3.41. Areas of low static pressure for case 6 of Speed Study: 50% Span ....................... 112 Figure 3.42. Zweifel coefficient vs. Efficiency for Blade Length Study .................................... 116 Figure 3.43. Zweifel Coefficient vs. Flow deviation at trailing edge for Blade Length Study .. 117 Figure 3.44. Difference of the peak efficiency and efficiency of the design vs. the Zweifel coefficient ................................................................................................................................... 118 Figure 3.45. Relative velocity vectors for case 1 of Blade Length Study: Span 50% ................ 119 Figure 3.46. Relative velocity vectors for case 2 of Blade Length Study: Span 50% ................ 119 Figure 3.47. Relative velocity vectors for case 3 of Blade Length Study: Span 50% ................ 120 Figure 3.48. Relative velocity vectors for case 4 of Blade Length Study: Span 50% ................ 120 Figure 3.39. Relative velocity vectors for case 5 of Blade Length Study: Span 50% ................ 120 Figure 3.40. Relative velocity vectors for case 6 of Blade Length Study: Span 50% ................ 121 Figure 3.41. Relative velocity vectors at leading edge for case 5 of Blade Length Study (Span 50%) ............................................................................................................................................ 121 Figure 3.42. Relative velocity vectors at leading edge for case 3 of Blade Length Study (Span 50%) ............................................................................................................................................ 122 Figure 3.43. Relative velocity vectors at leading edge for case 1 of Blade Length Study (Span 50%) ............................................................................................................................................ 122 Figure 3.44. Relative velocity vectors at trailing edge for case 5 of Blade Length Study (Span 50%) ............................................................................................................................................ 122 Figure 3.45. Relative velocity vectors at trailing edge for case 3 of Blade Length Study (Span 50%) ............................................................................................................................................ 123 Figure 3.46. Relative velocity vectors at trailing edge for case 1 of Blade Length Study (Span 50%) ............................................................................................................................................ 123 Figure 3.47. Relative velocity near trailing edge of case 3 and 5 of Blade Length Study.......... 124 Figure 3.48. Turbulence kinetic energy contour for cases 1j5 and 1j8: Span 50% .................... 125 Figure 3.49. Turbulence kinetic energy contour for designs 3n and 3n6: Span 50% ................. 125 Figure 3.50. Turbulence kinetic energy contour for designs 5n and 5n6: Span 50% ................. 125 Figure 3.51. Zweifel Coefficeint vs Exit Loss Coefficient for Blade Length Study .................. 126 Figure 3.52. Blade loading diagrams for case 3 of Blade Length Study: Span 50% .................. 128 Figure 3.53. Static Pressure Distributions for Case 3 of Blade Length Study (Span 50%) ........ 128 Figure 3.54. Blade loading diagrams for case 5 of Blade Length Study: Span 50% .................. 129 Figure 3.55. Static Pressure Distributions for Case 5 of Blade Length Study (Span 50%) ........ 129 Figure 3.56.Velocity triangles superimposed over velocity vector contours for design 2l of the Blade Length Study at leading and trailing edge (span 50%) ..................................................... 130 ix Figure 3.57.Velocity triangles superimposed over velocity vector contours for design 2l8 of the Blade Length Study at leading and trailing edges (span 50%) ................................................... 131 Figure 3.58. Relative velocity of case 6 for Blade Length Study at the leading edge along span ..................................................................................................................................................... 132 Figure 3.59. Relative velocity of case 6 for Blade Length Study at the trailing edge along span ..................................................................................................................................................... 133 Figure 3.60. Zweifel coefficeint vs Profile loss coefficient for Blade Length Study ................. 134 Figure 3.61. Average relative velocity from inlet to outlet for case 1 for Blade Length Study . 135 Figure 3.62. Average relative velocity from inlet to outlet for case 6 for Blade Length Study . 135 Figure 3.63. Absolute circumferential velocity for case 6 of the Blade Length Study at the trailing edge along span .............................................................................................................. 136 Figure 3.64. Absolute circumferential velocity for case 1 of the Blade Length Study at the trailing edge along span .............................................................................................................. 137 Figure 3.65. Average Absolute circumferential velocity from inlet to outlet of case 6 of Blade Length Study ............................................................................................................................... 138 Figure 3.66. Average Absolute circumferential velocity from inlet to outlet of case 1 of Blade Length Study ............................................................................................................................... 138 Figure 3.67. 𝑈𝜃𝐶𝑢 for case 1 of the Blade Length Study at the trailing edge across span ........ 139 Figure 3.68. Areas of low static pressure for case 1 of Blade Length Study (Span 50%) .......... 140 Figure 3.69. Areas of low static pressure for case 2 of Blade Length Study (Span 50%) .......... 140 Figure 3.70. Areas of low static pressure for case 3 of Blade Length Study (Span 50%) .......... 140 Figure 3.71. Areas of low static pressure for case 4 of Blade Length Study (Span 50%) .......... 141 Figure 3.72. Areas of low static pressure for case 5 of Blade Length Study (Span 50%) .......... 141 Figure 3.73. Areas of low static pressure for case 6 of Blade Length Study (Span 50%) .......... 141 x KEY TO SYMBOLS 𝐴𝑚 = meridonial cross-sectional area (𝑚2 ) 𝛼 = Absolute flow angle (𝑑𝑒𝑔𝑟𝑒𝑒𝑠) 𝑏 = Axial blade length (𝑚) 𝛽 = Relative flow angle(𝑑𝑒𝑔𝑟𝑒𝑒𝑠) 𝛽𝑏 = Blade angle (𝑑𝑒𝑔𝑟𝑒𝑒𝑠) 𝛽𝑚 = Mean relative flow angle (𝑑𝑒𝑔𝑟𝑒𝑒𝑠) 𝑚 𝐶 = Absolute flow speed ( 𝑠 ) ⃗⃗⃗ = Absolute flow speed (vector form) (𝑚) 𝐶 𝑠 𝑐 = Chord length (𝑚) 𝑚 𝐶𝑚 = Absolute flow speed in meridional direction ( 𝑠 ) 𝑚 𝐶𝑢 = Absolute flow speed in tangential direction ( 𝑠 ) 𝐶𝑌 = Tangential force coefficient 𝐶𝑑 =Drag coefficient 𝐶𝑥 =Axial force coefficient 𝐶𝑙 = Lift force coefficient 𝐷 = Outside diameter (𝑚) 𝐷𝑠 = Specific Diameter 𝑚2 𝑒̃ = Mass-specific shaft work output ( 𝑠2 ) 𝑚2 𝑒 = Total mass-specific work done on the fluid from leading edge to trailing edge ( 𝑠2 ) 𝑓 = friction factor xi 𝐹𝑥 = Axial force ( 𝐹𝑑 = Drag force ( 𝐹𝑙 = Lift force ( 𝑘𝑔 𝑚 𝑠2 𝑘𝑔 𝑚 𝑠2 𝑘𝑔 𝑚 𝑠2 ) ) ) 𝑚2 𝐹𝑓 = friction work lost per unit mass ( 𝑠2 ) 𝑚 𝑔 = gravitational acceleration constant (𝑠2 ) ℎ𝑏 = Blade height (𝑚) 𝑚2 ℎ = Enthalpy ( 𝑠2 ) 𝑚2 ℎ𝑜 = Stagnation enthalpy( 𝑠2 ) 𝐻 = total head (𝑚) 𝑖 ′ = Inlet incidence (𝑑𝑒𝑔𝑟𝑒𝑒𝑠) 𝑚2 𝑘 = Turbulence kinetic energy ( 𝑠2 ) 𝐾 = Friction loss coefficient from 90 degree turn 𝐿𝑡 = length of penstock pipe (𝑚) 𝑘𝑔 𝑚̇ =Mass flow rate ( 𝑠 ) 𝑘𝑔 ∆𝑚̇ = Mass flow rate through a single blade passage ( 𝑠 ) 𝑁 = Rotational speed ( 𝑟𝑜𝑡 𝑠 ) 𝑁𝑠 = Specific Speed (𝑟𝑜𝑡) 𝑁𝑟𝑝𝑚 = rotational speed (𝑟𝑜𝑡/𝑚𝑖𝑛) 𝑁𝑏 = number of blades 𝑘𝑔 𝑃 = Static pressure (𝑚𝑠2 ) xii 𝑘𝑔 𝑃𝑡 = Total Pressure (𝑚𝑠2 ) 𝑘𝑔 𝑃𝑎𝑡𝑚 = Atmospheric Pressure (𝑚𝑠2 ) 2 𝑚 𝑘𝑔 𝑄̇ = Heat input ( 𝑠3 ) 𝑅𝑒 = Reynolds Number 𝑟ℎ = Inside (hub) radius (𝑚) 𝑟𝑠 = Outside (shroud) radius (𝑚) 𝑟𝜃 = Distance from centerline in tangential direction (𝑚) 𝑟𝑚 = Mean radius (𝑚) 𝑟𝑏 = Radius of the penstock pipe (𝑚) 𝑘𝑔 𝜌 = Density (𝑚3 ) 𝑠 = Blade spacing (𝑚) 𝑡 = Blade thickness (𝑚) 𝑇 = Local temperature (𝐶𝑒𝑙𝑐𝑖𝑢𝑠) 𝑚 𝑢 = Local flow velocity ( 𝑠 ) 𝑚 𝑈 = Rotor tip speed( 𝑠 ) 𝑚 ⃗⃗⃗ 𝑈 =Absolute rotor velocity (vector form) ( 𝑠 ) 𝑚 𝑈𝜃 = Absolute speed of rotor at position along span ( ) 𝑠 3 𝑚 𝑉̇ = Volume flow rate ( 𝑠 ) 𝜔 = Angular velocity ( 𝑟𝑎𝑑 𝑠 ) 𝑚 𝑊 = Relative flow speed ( 𝑠 ) xiii 𝑚 ⃗⃗⃗⃗⃗ 𝑊 = Relative flow speed (vector form) ( 𝑠 ) 2 𝑚 𝑘𝑔 𝑊̇ = Shaft power output ( 𝑠3 ) 𝑚 𝑊𝑚𝑒𝑎𝑛 = Mean relative velocity ( 𝑠 ) 𝑚 𝑊𝑚 = Relative flow speed in meridional direction( 𝑠 ) 𝑚 𝑊𝑢 = Relative flow speed in tangential direction ( 𝑠 ) 𝑌 = Tangential blade force ( 𝑘𝑔 𝑚 𝑌𝑖𝑑 = Idealized blade force ( 𝑠2 ) 𝑘𝑔 𝑚 𝑠2 ) 𝑍 = Zweifel Coefficient 𝑧 = Position above datum (𝑚) 𝜉 = Loss coefficient 𝛿 ′ = Outlet deviation (𝑑𝑒𝑔𝑟𝑒𝑒𝑠) 𝜀𝑟 = surface roughness(𝑚) 𝜙 = Flow Coefficient 𝜓 = Blade loading coefficient 𝜓′ = Head coefficient 𝜙 ′ = Capacity Coefficient 𝜇 = Kinematic viscosity 𝜇𝑡 = Kinematic turbulence viscosity 𝛿𝑗𝑘 = Kroneker delta Ω𝑠 = Specific Speed (𝑟𝑎𝑑) 𝛾 = Hub-tip ratio xiv 𝜂 = Isentropic turbine efficiency 𝜂𝑡 = Total turbine efficiency 𝑘𝑔 𝜏𝑗𝑘 ′ = Reynolds stresses(𝑚𝑠2 ) 𝜏𝑎 = Time rate of change of momentum about a 𝑘𝑔 𝜏 = Shear stress(𝑚𝑠2 ) Subscripts 𝑚𝑎𝑥 = At 100% efficiency, idealized 1 = Turbine leading edge 2 = Turbine trailing edge a = Position a: water surface/inlet of penstock pipe b = Position b: outlet of penstock pipe/inlet of turbine entry annulus 1= Position 1: Turbine leading edge 2 = Position 2: Turbine trailing edge c = Position c: Outlet of turbine exit annulus, water level 𝑥 = axial direction ℎ = hub s = shroud i = ideal property (assuming 100% efficiency) k = index notation j = index notation 𝜃 = Tangential direction xv CHAPTER 1. INTRODUCTION AND BACKGROUND 1.1. Introduction to Hydroelectric Power Hydroelectricity is known as a renewable, sustainable, clean energy source with large potential for new development in the United States [1]. Hydroelectric power plants convert the kinetic and potential energy of water into electric power. To accomplish this, hydroelectric power plants use a turbine which converts the energy of the water into mechanical work, which is then transferred to a generator converting that mechanical work into electric power. Hydraulic (hydro) turbines usually have a stator which directs the water flow, as well as a runner or rotor. The stator directs the flow in a desired direction, using vanes or a nozzle. The rotor has blades which change the angular momentum of the flow, exerting a torque on the rotor inducing rotation [2]. The two main parameters which govern the amount of power which can be generated by a hydraulic turbine are the total available head 𝐻 and the volume flow rate 𝑉̇ . Head is defined as the height difference between two water surfaces. Most hydropower plants use a dam to generate the head necessary for hydroelectric power generation. These dams hold back water, creating a reservoir with a high water level on one side of the dam. The potential energy of the water behind the dam is used to extract energy. Hydroelectric dams often use an intake pipe, called a penstock, to feed water into the turbine. After the water flows through the turbine, it flows through an outlet tube called a draft tube which has an expanding cross-section. The draft tube is designed to decelerate the water. The use of a draft tube also allows for the turbine to be placed above the lower water level and extract the full potential energy of the water. Therefore the draft tube ensures that the head of the water below the tail race level [3]. 1 An example of this is shown below in Figure 1.1, a diagram of a dam with a hydroelectric system. Figure 1.1. Hydroelectric Dam Schematic The amount of electric power generated by a hydroelectric system (𝑊̇ ) depends on the total efficiency of the system (𝜂𝑡 ), the total head 𝐻, the gravitational acceleration constant 𝑔, and the volume flow rate 𝑉̇ , and is calculated using equation 1-1 [4]: 𝑊̇ = 𝑔𝑉̇ 𝐻𝜂𝑡 (1-1) The total efficiency of the system depends on the isentropic efficiency of the turbine rotor, the amount of friction losses in the penstock and draft tube, as well as the efficiency of the generator, among other factors. 1.2. Hydroelectric Power Challenges Hydroelectric sites can be categorized by the potential for power output. They are divided into Large, Medium, Small, Mini, and Micro categories. This thesis focuses primarily on Mini hydro (between 100 and 1000 kW) and Micro hydro (up to 100 kW) turbine designs and development. Up to this point, constraints and problems related to the development of mini/micro hydropower include [5]: 1. High capital cost of hydropower generation equipment 2 2. High cost of civil works typically associated with mini/micro hydropower 3. High operations/maintenance cost The US Department of Energy has suggested pre-packaged, pre-assembled modular low cost hydroelectric units are needed to solve this problem [1]. It has been suggested that a reduction in the capital cost of mini/micro hydroelectric units could lead to additional growth of hydropower generation in the United States [1]. This thesis focuses on the first problem, specifically the high capital cost of the turbine equipment. 1.3. Hydroelectric Dam Retrofits In order to achieve the goal of expanding hydropower use, it is advantageous to look for potential sites with the lowest potential cost of installation and civil works. “Hydroelectric dam retrofits” are considered to be a cost-effective hydropower development strategy. Hydroelectric dam retrofits involve installing hydropower generation equipment into existing dams that have the necessary water volume flow and head for potential power generation [6]. The power generated could either be used to supplement the grid or offset electricity costs for municipal dam owners. In a report prepared for the US Department of Energy, The Oak Ridge National Laboratory stated that the primary advantage of hydroelectric dam retrofits compared to new hydropower development is that “many of the costs and environmental impacts of dam construction have already been incurred at NPDs (Non-Powered-Dams) and may not be significantly increased by the incorporation of new energy production facilities. Thus, the development of some NPD’s for energy purposes is assumed to be achievable with lower installed cost, lower levelized cost-of-energy, fewer barriers to development, less technological and business risk, and in a shorter time frame than development requiring new dam construction” 3 [6]. The Oak Ridge National Laboratory has estimated that a potential additional 12.6 gigawatts of power could be generated from hydroelectric dam retrofits in the United States. It is considered an abundant potential source of renewable energy in the US. There are over 80,000 non-powered dams in the United States, and it is estimated 54,000 of these have the water flow and head necessary for considerable power output if a hydroelectric system was installed [6]. Civil works contribute around 40% of the total cost of conventional small hydro projects [7]. The civil work projects that need to be done for a hydroelectric dam retrofit include the intake, penstock, and power house construction. Civil works cost also include any temporary infrastructure needed during installation and construction. Typically when installing new hydropower equipment into an existing dam, the dam must be de-watered and a new temporary dam must be constructed. The cost of the civil works projects associated with a hydroelectric dam retrofit are highly dependent on the design of the hydroelectric system selected. It is advantageous to design the hydroelectric dam retrofit system such that extensive modification of the existing dam structure is not required. Many non-powered dams in the United States have spillways from which water flows. Most spillways have gates which control the rate at which water flows through the dam. Figure 1.2 below shows the Bellaire Dam in Antrim County, Michigan, and Figure 1.3 shows a spillway gate at the dam. 4 Figure 1.2. Bellaire Dam, Antrim County, Michigan Figure 1.3. Spillway at the Bellaire Dam, Antrim County, Michigan The analytical model in this work was set up to obtain CFD boundary conditions which were applicable for a proposed application of a hydraulic turbine for use as a retrofit turbine unit to fit on the front of dam spillways. The setup analyzed in this work is illustrated in Figure 1.4 below. 5 Spillway Turbine Figure 1.4. Hydroelectric dam retrofit schematic 1.4. Selection of Type of Water Turbine When selecting a turbine system to be used for a potential site for hydroelectric plant installation, a number of different factors are considered. The first being the volume flow rate, head, and power obtainable given the geographic constraints of the hydropower site. The degree of complexity involved to build/install the hydroelectric system, as well as maintenance requirements of the turbine technology being evaluated are also considered. The portability and shipping cost of the system are factors as well as the degree of modularity of the turbine being considered for selection and potential environmental impacts. Modular turbine systems are of value as it allows for the turbine to be broken up into smaller components for easy maintenance and replacement in the field [2]. Before going forward on a hydropower project, the capital and operating costs of the system are weighed against the value of the energy the plant will generate. Engineers often pick the type of turbine to be used for a particular application by observing the designed flow rate and head, and picking the type of turbine based off of recommendations of literature. 6 1.5. Woven Wheel Hydro Turbine A technology invented by Dr. Norbert Mueller at Michigan State University known as the Woven Wheel offers promise to reduce the capital cost and complexity of hydraulic turbines. The Woven Wheel is a manufacturing process which involves manufacturing turbomachinery blades by winding the wheel out of continuous composite fiber strands. The wheels are designed such that the fibers are wound in tension. For the Woven Wheel manufacturing technique, the A technology invented by Dr. Norbert Mueller at Michigan State University known as the Woven Wheel offers promise to reduce the capital cost and complexity of hydraulic turbines. The Woven Wheel is a manufacturing process which involves manufacturing turbomachinery blades by winding the wheel out of continuous composite fiber strands. The wheels are designed such that the fibers are wound in tension. For the Woven Wheel manufacturing technique, the continuous fiber is wound around a low-cost mandrel which can be 3-D printed. The technology allows for motor and generator components to be integrated directly into the hub or shroud, decreasing the cost of producing modular turbines or compressors which use the Woven Wheel as the rotor [8]. The winding of these wheels can be done by hand or using a commercially available computer controlled winding machine. The Woven Wheel design includes an outer shroud [9]. This winding process allows for a large number of different winding patterns to be employed. Figure 1.5(a) below shows a computer rendering of one such winding pattern. Figure 1.5(b) shows the wound turbomachine wheel inside of the mandrel, and Figure 1.5(c) shows the wheel after removal from the mandrel. 7 Figure 1.5. (a) 3-D model of Woven Wheel winding scheme (b) Woven Wheel in mandrel (c) Woven wheel removed from mandrel This technology developed at Michigan State University allows for lower cost manufacturing of axial water turbines with integrated generator components. Inclusion of an integrated generator along the shroud of the rotor can decrease the part count and required maintenance as well. A prototype Woven Wheel water turbine was successfully tested for a tidal turbine application in a tow tank at the Marine Hydrodynamic Lab, at the University of Michigan, shown below in Figure 1.6. Figure 1.6. Woven Wheel tidal turbine being tested inside the tow tank Integrated motor components have been demonstrated in the Woven Wheel’s application as a compressor; however, the integration of the generator components in the Woven Wheel’s application as a water turbine has yet to be prototyped. The turbine rotor tested, manufactured using the Woven Wheel manufacturing method was wound from Kevlar-49. Kevlar fiber has a specific strength three times higher than titanium alloys [10]. The advantages of using this Woven Wheel technology as a water turbine are discussed further below. 8 1.5.1. Lightweight, Modular Design The Woven Wheel modular turbine or compressor units can be assembled in a row to form a modular multistage unit. Kevlar is a high strength, low cost, low weight fiber material which can be considered a good choice of composite material for winding a Woven Wheel for use as a water turbine. Figure 1.7 below shows a rendering of one such modular design. A recent Woven Wheel prototype weighed approximately 50% of a wheel of the similar geometry machined from 6061 T6 aluminum. Figure 1.7. Modular Woven Wheel compressor unit design 1.5.2. Low-cost Additive Manufacturing The Woven Wheel can be manufactured by winding a continuous carbon fiber bundle around a 3-D printed mandrel to form the shape of the blades. The wheel with mandrel is then infused with epoxy and cured, after which the wheel is removed from the mandrel. Preliminary prototyping of manufacturing Woven Wheels using this method has taken place at Michigan State University. The cost of manufacturing traditional turbine blades with an integrated shroud with traditional turbine manufacturing methods is high as it would require more raw material, and machining processes. The weight of the turbine would inevitably increase as well. Figure 1.8 below shows a 3-D printed mandrel and the winding process. Although the process requires 9 further maturation, prototypes have demonstrated it costs less to purchase the materials and produce compared to traditional rotors machined out of metal. Figure 1.8. (a) 3-D printed mandrel. (b) carbon fiber winding. (c) mandrel post-winding after being dipped into epoxy/resin and cured. (d) Woven Wheel after removal from mandrel. 1.5.3. Rapid Customization and Manufacturing Although this manufacturing process is still being refined and perfected, this additive manufacturing process allows for the time between design and production to be minimal. New designs derived from analytical models can be tested in CFD, a mandrel can be manufactured and wound in a matter of hours. This allows for faster, lower cost production and rapid customization. This added versatility lends itself well to the application of the technology for use as a water turbine. Individual hydroelectric applications are unique and are often custom 10 projects, and the ability to adjust the Woven Wheel manufacturing technique to fit any required size is advantageous. Currently only blade profiles with constant thickness can be manufactured using the Woven Wheel method. The turbine designs investigated in this thesis all have constant blade thickness from hub to shroud, and leading to trailing edge. 1.5.4. Outer Shroud Eliminates Tip Leakage and Reduces Fish Mortality Fishes can suffer injuries and death passing through hydraulic turbines. This is considered a significant environmental concern when considering any new hydroelectric development [11]. Fish injuries and death while passing through hydraulic turbines is a result of rapid pressure changes along flow path, cavitation, narrow gaps between rotating parts and stationary structures, and fish collision with structures including turbine runner blades and guide vanes. Traditional axial-flow turbines (such as Kaplan turbines) are not shrouded, there is a gap between the turbine runner and outer housing which is the most common source of the mortality of fish passing through hydraulic turbines [12]. Turbine manufactures like Voith have attempted to solve this problem by implementing design changes which reduce the factors listed above. Voith has developed a “Minimum Gap Runner” technology which features specially contoured and machined runner blades to minimize the gaps between the turbine runner blades and the hub, as well as the stationary outer housing. The minimization of the gap between turbine runner blades and the housing also reduces tip leakage, improving performance [13]. Computational Fluid Dynamics (CFD) is often used to ensure that cavitation, shear stresses, and rapid pressure changes are kept to a minimum. Gaps between the turbine rotor blades and outer housing can lead to fish death. Companies like Voith have developed technologies for axial hydraulic turbines which reduce the 11 size of this gap, however almost all axial hydraulic turbines do not have this feature [13]. Using the Woven Wheel as a hydraulic turbine reduces the risk of fish death as a result of the gap between rotor blades and the outer housing as the shroud is integrated into the design of the blades. This also reduces losses associated with tip leakage. 1.6. Definition of Turbine Type Investigated This work explores the viability of using an axial turbine with simplified geometry for low head hydroelectric applications. Constraints were placed on the design to keep manufacturing costs as low as possible, and to allow the Woven Wheel method to be employed for manufacturing the turbine rotor. The turbines analyzed in this work have the following features:  Axial flow  Circular arc blade profile  Constant thickness blade profile  Constant OD/ID  Shrouded rotor  No inlet or outlet guide vane  No-pre swirl, 𝐶𝑢1 = 0 The performance of the hydro turbine designs tested in CFD could be improved by employing blade profiles, adjusting the OD/ID along the flow path axially, but these were not included in the designs, this was done to keep manufacturing production cost low and to make the designs applicable to the Woven Wheel manufacturing method. In this work turbines of this type referred to as CTPAT, or Constant Thickness Profile Axial Turbine. 12 The work in this thesis applies for turbine designs with shrouded rotors manufactured from metal as well as Woven Wheels. The CTPAT scheme is applicable with the modular Woven Wheel turbine concept which has been proposed. The use of constant thickness blade profiles makes the CTPAT scheme different from those turbines most commonly employed in the field. The analytical model described in section 2.3 was set up to obtain CFD boundary conditions which were applicable for a proposed application of a CTPAT turbine for use as a retrofit turbine unit to fit on the front of dam spillways. 1.7. Conservation of Energy The first law of thermodynamics is applied through a control volume to obtain an equation for the steady flow energy balance for a turbine. A control volume representing a turbine is illustrated in Figure 1.9 below. For the analyses in this work, “1” represents the properties at the leading edge of the turbine, and “2” represents the properties at the trailing edge of the turbine. Equation 1-2 below shows the energy balance, where 𝑄̇ is the heat transfer from the surroundings to the control volume measured in Joules per second, ℎ is mass-specific enthalpy, 𝐶 is absolute flow speed, 𝑧 represents the elevation above the datum, and 𝑊̇ represents the power that is transferred from the fluid to the blades of the turbomachine via the shaft (shaft power). 13 Figure 1.9. Turbomachine control volume 1 𝑄̇ − 𝑊̇ = 𝑚̇ [(ℎ2 − ℎ1 ) + (𝐶22 − 𝐶12 ) + 𝑔(𝑧2 − 𝑧1 )] 2 (1-2) The heat transfer from the surroundings to the control volume 𝑄̇ is negligible for water turbines, and is assumed to be equal to zero. To derive an expression for the isentropic efficiency of the turbine, equation 1-2 is rewritten in differential form below in equation 1-3 [4]: 1 𝑑𝑊̇ = 𝑚̇ [𝑑ℎ + 𝑑(𝑐 2 ) + 𝑔𝑑𝑧)] 2 (1-3) ̇ is the total energy extraction from the fluid per second, and is the maximum 𝑊𝑚𝑎𝑥 possible power output of the turbine. For an isentropic process, 𝑑ℎ = 𝑑𝑝 𝜌 ̇ can then be [4]. 𝑊𝑚𝑎𝑥 expressed using equation 1-4: ̇ = 𝑚̇ [ 𝑊𝑚𝑎𝑥 (𝑃2 − 𝑃1 ) 1 2 + (𝐶2 − 𝐶12 ) + 𝑔(𝑧2 − 𝑧1 )] 𝜌 2 (1-4) ̇ can For incompressible flow, the total energy extraction from the fluid per second 𝑊𝑚𝑎𝑥 be rewritten in terms of the total available head 𝐻, which takes into account gravitational potential energy, kinetic energy, and enthalpy difference, seen in equation 1-3, applied to the energy equation in equation 1-5 below [4]: 14 𝑔𝐻 = 𝑃 1 2 + 𝐶 + 𝑔𝑧 𝜌 2 (1-5) ̇ , equation The expression for the total energy extraction from the fluid per second, 𝑊𝑚𝑎𝑥 () can then be rewritten in terms of 𝐻, shown below in equation 1-6: ̇ = 𝑚̇𝑔(𝐻1 − 𝐻2 ) 𝑊𝑚𝑎𝑥 (1-6) The isentropic turbine efficiency 𝜂 is defined as the ratio of the shaft power output 𝑊̇ to ̇ , and defined below in equation 1-7 the total energy extraction from the fluid per second 𝑊𝑚𝑎𝑥 [4]. Isentropic turbine efficiency is referred to as “efficiency” in this work. 𝜂= 𝑊̇ 𝑊̇ = ̇ 𝑚̇𝑔(𝐻1 − 𝐻2 ) 𝑊𝑚𝑎𝑥 (1-7) The mass-specific form of the shaft power output 𝑊̇ is denoted by 𝑒̃ , measured in meters squared per seconds squared, 𝑒̃ is defined in equation 1-8 below: 𝑒̃ = 𝑊̇ 𝑚̇ (1-8) Using equation 1-8, the expression for efficiency in equation 1-7 can then be rewritten in terms of the mass-specific shaft work 𝑒̃ , shown below in equation 1-9: 𝜂= 𝑒̃ 𝑔(𝐻1 − 𝐻2 ) (1-9) ̇ . 𝑒 denotes the mass-specific A variable is defined, 𝑒, the mass-specific form of 𝑊𝑚𝑎𝑥 work extracted from the fluid from the leading to trailing edge, which includes losses in the turbine, and is defined below in equation 1-10: 𝑒= ̇ 𝑊𝑚𝑎𝑥 𝑒̃ = = 𝑔(𝐻1 − 𝐻2 ) 𝑚̇ 𝜂 15 (1-10) Work extraction from a hydraulic turbine can be expressed in terms of the total (stagnation) pressure, 𝑃𝑡 . 𝑃𝑡 can be expressed in terms of static pressure, dynamic pressure, and the gravitational head, shown in equation 1-11 below [4]: 𝑃𝑡 = 𝑃 1 2 + 𝐶 + 𝑔𝑧 𝜌 2 (1-11) The mass-specific energy extracted from the fluid from leading to trailing edge, 𝑒, is expressed in terms of the total pressure 𝑃𝑡 at leading and trailing edges in equation 1-12 below for incompressible flow for a hydraulic turbine: 𝑃𝑡2 = 𝑃𝑡1 − 𝑒𝜌 (1-12) A portion of the total pressure drop across a turbine stage (𝑃𝑡1 − 𝑃𝑡2 ) is converted into useful shaft work (𝑒̃ 𝜌). The portion of the pressure drop that is not converted into shaft work are considered losses. The losses can be expressed using a reordered version of equation 1-12, in terms of the total loss coefficient 𝜉 and the absolute velocity at the trailing edge, 𝐶2 , shown below in equation 1-13. 1 𝑃𝑡1 − 𝑃𝑡2 = 𝑒̃ 𝜌 + 𝜌𝜉𝐶22 2 (1-13) Total loss coefficient 𝜉 can be expressed in terms of its components, a sum of the sources of loss in a turbine stage, where 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 represents losses in the blade row, 𝜉𝑒𝑥𝑖𝑡 represents losses between the trailing edge and the outlet, 𝜉𝑒𝑛𝑡𝑟𝑦 represents losses between the turbine inlet and leading edge, expressed below in equation 1-14: 𝜉 = 𝜉𝑒𝑛𝑡𝑟𝑦 + 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 + 𝜉𝑒𝑥𝑖𝑡 (1-14) The shaft power output of the turbine 𝑊̇ can be calculated using the mass-specific shaft work 𝑒̃ and the mass flow rate 𝑚̇, shown below in equation 1-15: 16 𝑊̇ = 𝑒̃ 𝑚̇ (1-15) 1.8. Definitions of Geometry and Design Parameters Figure 1.10. Example velocity triangle The variables used to generate a turbine rotor design for CFD analysis are defined in this ⃗ . The absolute velocity of the fluid is 𝐶 , and section. Absolute velocity of the turbine rotor is 𝑈 ⃗⃗⃗ , defined below in equation 1-16. the velocity of the fluid relative to the rotor is 𝑊 ⃗⃗⃗ = 𝐶 − 𝑈 ⃗ 𝑊 (1-16) ⃗⃗⃗ are often expressed in terms of their meridonial and tangential The vectors 𝐶 and 𝑊 (wheel rotation direction) components. 𝐶𝑚 is the absolute velocity of the fluid in the meridonial direction, and 𝑊𝑚 is the flow speed in the meridional direction relative to the wheel. 𝐶𝑢 is the absolute velocity of the fluid in the direction of the rotation of the wheel, and 𝑊𝑢 is the flow speed in the rotation direction of the wheel relative to the wheel. Velocity triangles help to visualize how the velocities relate to each other. Figure 1.10 above shows an example velocity 17 triangle. The magnitude of the absolute flow velocity, 𝐶, is calculated using the meridonial and tangential velocity components shown below in equation 1-17: 2 𝐶 2 = 𝐶𝑢2 + 𝐶𝑚 (1-17) Similarly the magnitude of the relative flow velocity 𝑊 can be calculated using the meridional and tangential relative velocity components, seen in equation 1-18: 𝑊 2 = 𝑊𝑢2 + 𝑊𝑚2 (1-18) ⃗ vector, and Absolute flow angle 𝛼 is defined as the angle between the 𝐶 vector and the 𝑈 can be calculated using the components of the 𝐶 vector, expressed below in equation 1-19: 𝛼 = cos −1 𝐶𝑢 𝐶𝑚 = = sin−1 ( ) 𝐶 𝐶 (1-19) ⃗⃗⃗ vector and the Similarly the relative flow angle, 𝛽, is defined as the angle between the 𝑊 ⃗ vector, and can be calculated using the components of the 𝑊 ⃗⃗⃗ vector, expressed below in 𝑈 equation 1-20: β = sin−1 𝑊𝑚 𝑊𝑢 = cos −1( ) 𝑊 𝑊 (1-20) Figure 1.11 below shows an example CTPAT axial turbine geometry, where ℎ𝑏 is the blade height, 𝑟ℎ is the inner (hub) radius, and 𝑟𝑠 is the outside (shroud) radius of the turbine rotor, and 𝜔 is the angular velocity of the rotor. Position 1 represents the leading edge, position 2 represents the trailing edge. 18 Figure 1.11. Axial turbine geometry parameters For the analysis in this work, the outside (shroud) radius of the rotor, 𝑟𝑠 , and the rotational speed of the rotor in RPM are all input. The hub-tip ratio 𝛾 is input as well. The inner (hub) radius 𝑟ℎ can be calculated using the shroud radius and the hub-tip ratio, shown in equation 1-21 below: 𝑟ℎ = 𝑟𝑠 𝛾 (1-21) Angular velocity 𝜔 can be calculated using the rotational speed given in rotations per minute 𝑁𝑟𝑝𝑚 , which was input for this analysis, using equation 1-22,then the tip speed of the wheel (𝑈) can then be calculated using equation 1-23. 𝜔 = 2𝜋 𝑁𝑟𝑝𝑚 60 𝑈 = 𝑟𝑠 𝜔 (1-22) (1-23) The tangential speed of the rotor for a position along the span is denoted by 𝑈𝜃 , where “𝜃” represents a position along the span, calculated using equation 1-24 below, where 𝑟𝜃 is the radial distance from the centerline to the position along the span being considered. This is 19 different than the tip speed 𝑈 in that tip speed uses outside radius. For example, the tangential speed of the rotor at the inner radius/hub is 𝑈ℎ = 𝑟ℎ 𝜔, calculated using the inner radius, 𝑟ℎ . 𝑈𝜃 = 𝑟𝜃 𝜔 (1-24) The inner radius and outside radius of each turbine in this analysis were kept constant form leading to trailing edges. Thus the tip speed of the wheel is constant from the leading to trailing edges, so 𝑈 = 𝑈1 = 𝑈2 . For this analysis the thickness of the blades (𝑡) and the number of blades (𝑁𝑏 ) were also input. The thickness of the blades is constant from leading to trailing edge, and is also constant across the blade span. Blade height ℎ𝑏 is calculated for the IGV and rotor geometries using equation 1-25: ℎ𝑏 = 𝑟𝑠 − 𝑟ℎ (1-25) Meridional cross-sectional area 𝐴𝑚 is then calculated for the axial rotor using equation 1-26: 𝐴𝑚 = 𝜋(𝑟𝑠2 − 𝑟ℎ2 ) − 𝑡ℎ𝑏 𝑁𝑏 (1-26) Mass flow rate 𝑚̇ can be calculated using the meridional cross-sectional area 𝐴𝑚 , density 𝜌, and absolute meridonial flow speed 𝐶𝑚 , shown below in equation 1-27: 𝑚̇ = 𝐶𝑚 𝐴𝑚 𝜌 (1-27) Equation 1-28 below is referred to as the equation of continuity, and states that mass flow rate, 𝑚̇, stays constant from inlet to outlet of the turbomachine control volume. 𝑚̇ stays constant from leading to trailing edge, and from inlet to outlet as there is no accumulation of water within the control volume. Equation 1-29 is a reordered version of equation 1-28, using the definition of mass flow rate in equation 1-27: 𝑚̇ 1 = 𝑚̇ 2 (1-28) 𝐶𝑚1 𝐴𝑚1 𝜌1 = 𝐶𝑚2 𝐴𝑚2 𝜌2 (1-29) 20 The mass flow rate of water flow (𝑚̇) as well as the density of the water (𝜌) are input. Using these input properties and known geometry, meridional absolute flow speeds along the flow path can be calculated using equation 1-30 below, after rearranging the definition of mass flow rate expressed in equation 1-27: 𝐶𝑚 = 𝑚̇ 𝐴𝑚 𝜌 (1-30) Flow speed in the direction of the wheel relative to the wheel 𝑊𝑢 is then calculated using ⃗⃗⃗ vector equation 1-16 ,in the tangential direction, seen below in equation the definition of the 𝑊 1-31: 𝑊𝑢 = 𝑈 − 𝐶𝑢 (1-31) Similarly by applying equation 1-16 in the meridional direction 𝑊𝑚 is calculated using equation 1-32 below: 𝑊𝑚 = 𝐶𝑚 (1-32) Figure 1.12 below illustrates an example CTPAT axial turbine blade cascade with velocity triangles at leading and trailing edges. Geometric parameters are defined, where 𝑏 is the axial blade length, 𝑠 is the spacing between blades, 𝑐 is the chord length, and 𝑡 is the blade thickness. Circular arc blade profiles are used for the analytical model and the CFD simulations. For this analysis geometric parameters 𝑡, 𝑁𝑏 , 𝑟𝑜 , 𝑟𝑖 , 𝑏 are each provided as an input. For the analyses in this work, “1” represents the properties at the leading edge, and “2” represents the properties at the trailing edge of the turbine. 21 Figure 1.12. Turbine blade cascade with velocity triangles Blade spacing 𝑠 can be calculated using equation 1-34 below, which uses known geometric inputs and mean radius 𝑟𝑚 , which is calculated using equation 1-33 below: (𝑟 2 + 𝑟𝑠2 ) 𝑟𝑚 = √ ℎ 2 𝑠= 2𝜋𝑟𝑚 𝑁𝑏 22 (1-32) (1-33) Flow does not follow the blade path perfectly, the difference between the ideal trailing edge relative flow angle if flow followed the blades perfectly and the actual trailing edge relative flow angle is referred to as the outlet deviation, 𝛿 ′ , defined in equation 1-34 below, where βb2 is the angle between the blade at the trailing edge and the ⃗⃗⃗ 𝑈 vector, and β2 is the relative flow angle at the trailing edge: 𝛿 ′ = β2 − βb2 (1-34) Similarly the incidence is the difference between the leading edge flow angle and the blade leading edge angle, shown below in equation 1-35: 𝑖 ′ = β1 − βb1 (1-35) The mass-specific shaft work 𝑒̃ can be related to the rotor speed and absolute tangential velocity using the Euler turbomachinery work equation, derived using conservation of angular momentum. Conservation of momentum equates the sum of external forces acting on a fluid to the rate of change of momentum. Applied to turbomachinery, the force applied by a fluid onto the blades is caused by the acceleration of the fluids passing through the blades [4]. Equation 136 below applies conservation of momentum in the x-direction for a control volume where fluid enters with uniform velocity 𝐶𝑥1 in the x direction and leaves with uniform velocity 𝐶𝑥2 in the x direction, 𝑚 represents the mass of a fluid element, and 𝐹𝑥 represents forces acted on the control volume in the x-direction [4]: ∑ 𝐹𝑥 = 𝑑 (𝑚𝐶𝑥 ) = 𝑚̇(𝐶𝑥2 − 𝐶𝑥1 ) 𝑑𝑡 (1-36) Conservation of momentum can be applied to relate change in tangential flow speed to the change in angular momentum and shaft power. The sum of all moments of all forces acting on a system about an axis is equal to the time rate of change of angular momentum about that axis [4]. This is seen below in equation 1-37, where 𝜏𝑎 is the time rate of change of angular 23 momentum about axis A, 𝑟 is the distance between the center of mass to the rotation axis A and 𝐶𝜃 is the component of velocity perpendicular to both the axis A and the radius vector: 𝜏𝑎 = 𝑚 𝑑 (𝑟𝐶𝜃 ) 𝑑𝑡 (1-37) Equation 1-37 can then be applied for an arbitrary turbomachine, with radius at leading edge 𝑟1, radius at trailing edge 𝑟2 , with tangential velocity at the leading and trailing edges 𝐶𝑢1 and 𝐶𝑢2 respectively for one dimensional steady flow, shown below in equation 1-38 [4]: 𝜏𝑎 = 𝑚̇(𝑟2 𝐶𝑢2 − 𝑟1 𝐶𝑢1 ) (1-38) 𝑈 Applying the definition of tip speed 𝑈 by plugging in 𝑟 = 𝜔 to equation 1-38, equation 139 is formed: 𝜏𝑎 𝜔 = 𝑚̇(𝑈2 𝐶𝑢2 − 𝑈1 𝐶𝑢1 ) (1-39) The product 𝜏𝑎 𝜔 is equal the shaft power output of the turbine 𝑊̇ . Using equation 1-8, equation 1-39 is rearranged in terms of the mass-specific shaft work 𝑒̃ , shown below in equation 1-40: 𝑒̃ = 𝑈2 𝐶𝑢2 − 𝑈1 𝐶𝑢1 (1-40) The above is a form of the Euler turbomachinery work equation, which is valid for steady, adiabatic flow [4]. For the analysis in this work, tip radius stays constant from leading to trailing edge such that 𝑈1 = 𝑈2 . Equation 1-40 can then be simplified to form equation 1-41: 𝑒̃ = 𝑈(𝐶𝑢2 − 𝐶𝑢1 ) (1-41) By rearranging equation 1-41 and applying equation 1-10, 𝑒̃ can then be expressed in terms of the total head 𝐻 difference from leading to trailing edge, shown below in equation 1-42: 𝑈(𝐶𝑢2 − 𝐶𝑢1 ) = 𝜂𝑔(𝐻1 − 𝐻2 ) 24 (1-42) 1.9. Turbine Cascade Forces An example pressure distribution across a turbine stage is shown below in Figure 1.13, where 𝑃𝑝 is the pressure curve on the pressure side of the blade, 𝑃𝑠 is the pressure curve on the suction side. Figure 1.13. Pressure distribution across a turbine stage The area between the 𝑃𝑝 and 𝑃𝑠 curves in Figure 1.13 is equal to the tangential force 𝑌 acting on the flow imparted by a single blade, calculated by integrating the pressure difference from the leading edge to the trailing edge, expressed below in equation 1-43. 1 𝑥 𝑌 = 𝑏 ∫ (𝑃𝑝 − 𝑃𝑠 ) 𝑑 ( ) 𝑏 0 (1-43) To derive an alternate form of equation 1-43, conservation of angular momentum is applied in the tangential direction, the tangential force can be expressed using equation 1-45 assuming ̇ is calculated, which represents the mass flow rate through a constant axial velocity. First ∆𝑚 single blade passage, using equation 1-44 [4]: 25 ̇ = 𝐶𝑚 𝑠ℎ𝑏 𝜌 ∆𝑚 (1-44) ̇ (𝑊𝑢2 − 𝑊𝑢1 ) 𝑌 = ∆𝑚 (1-45) Force in the axial direction 𝐹𝑥 can be derived using conservation of momentum, written below in equation 1-46, assuming axial velocity stays constant from leading to trailing edge: 𝐹𝑥 = (𝑃1 − 𝑃2 )𝑠ℎ𝑏 (1-46) These forces can be used to calculate lift and drag forces 𝐹𝑙 and 𝐹𝑑 on the blades. Figure 1.14 below shows an example turbine blade cascade with blade forces. Figure 1.14. Example turbine blade cascade with blade forces 26 First mean flow angle 𝛽𝑚 is calculated using equation 1-47, necessary for the calculation of lift and drag forces. 1 tan 𝛽𝑚 = (tan 𝛽1 + tan 𝛽2 ) 2 (1-47) Mean relative velocity 𝑊𝑚𝑒𝑎𝑛 can then be calculated using the mean relative flow angle 𝛽𝑚 and the relative meridional velocity 𝑊𝑚 , shown below in equation 1-48: 𝑊𝑚𝑒𝑎𝑛 = 𝑊𝑚 / sin 𝛽𝑚 (1-48) Lift and drag forces 𝐹𝑙 and 𝐹𝑑 can be calculated using equation 1-49 and equation 1-50 respectively, below: 𝐹𝑙 = 𝐹𝑥 cos 𝛽𝑚 + 𝑌 sin 𝛽𝑚 (1-49) 𝐹𝑑 = Ycos𝛽𝑚 − 𝐹𝑥 sin 𝛽𝑚 (1-50) Tangential force coefficient 𝐶𝑌 , axial force coefficient 𝐶𝑥 , lift coefficient 𝐶𝑙 , and drag coefficient are expressed in equations 1-51, 1-52, 1-53, and 1-54 respectively. They are useful non-dimensional coefficients which are a measure of the blade forces defined in equations 1-45, 𝜌 2 1-46, 1-49, and 1-50 respectively, compared to the mean dynamic pressure , 2 𝑊𝑚𝑒𝑎𝑛 , multiplied by the area on which it acts, 𝑠ℎ𝑏 . 𝑌 𝐶𝑌 = 𝜌 2 2 𝑊𝑚𝑒𝑎𝑛 𝑠ℎ𝑏 (1-51) 𝐹𝑥 𝐶𝑥 = 𝜌 2 𝑊 𝑠ℎ 2 𝑚𝑒𝑎𝑛 𝑏 (1-52) 𝐹𝑙 𝐶𝑙 = 𝜌 𝑊 2 𝑠ℎ 2 𝑚𝑒𝑎𝑛 𝑏 (1-53) 𝐹𝑑 𝐶𝑑 = 𝜌 2 𝑊 𝑠ℎ 2 𝑚𝑒𝑎𝑛 𝑏 (1-54) 27 1.10. Dimensionless Coefficients 1.10.1. Flow and Capacity Coefficients The flow coefficient, 𝜙, is a non-dimensional coefficient that depends on flow rate and rotational speed. The flow coefficient is used to compare different types of turbines. It relates the axial flow velocity to the speed of the blade, and is defined below in equation 1-55 [15]: 𝜙= 𝐶𝑚 𝑈 (1-55) The capacity coefficient, 𝜙 ′ , similar to the flow coefficient, is a non-dimensional parameter which is effected by the rotational speed, rotor diameter, and flow rate. The capacity coefficient is defined below in equation 1-56: 𝜙′ = 𝑉̇ 𝑁𝐷3 (1-56) 1.10.2. Blade Loading and Head Coefficients The blade loading coefficient 𝜓 provides a measure of the work extraction for a turbine stage. The blade loading coefficient relates the mass-specific shaft work 𝑒̃ to the blade tip speed 𝑈. is defined below in equation 1-57: [4]. 𝜓= 𝑒̃ 𝑈2 (1-57) The equation is rearranged for an adiabatic axial water turbine with constant shroud radius, plugging in equation 1-10 into equation 1-57 to obtain equation 1-58 below: [4] 𝜓= 𝑒̃ 𝜂𝑔𝐻 = 𝑈 2 𝜋 2 𝑁 2 𝑟𝑠2 (1-58) The head coefficient 𝜓′ is similar to the blade loading coefficient in that it relates the work extraction to the rotational speed. However, blade loading coefficient 𝜓 includes isentropic efficiency in the expression, while head coefficient 𝜓′ uses total work extraction from the fluid, 28 𝑒, in the expression. The head coefficient 𝜓′ is defined below for an axial water turbine with constant shroud radius using equation 1-59: 𝜓′ = 𝑒 𝑔𝐻 = 2 (𝑁𝐷) (𝑁𝐷)2 (1-59) Turbines which extract more work at a low tip speed have higher blade loading and head coefficients. 1.10.3. Specific Speed Specific speed is a non-dimensional quantity that is used to describe and categorize turbomachinery. Specific speed, also called shape parameter, is discussed by Horlock [15] as a useful tool to help select the type of machine that will give highest efficiency for a given application. The specific speed was derived by raising the head and capacity coefficients to a power, eliminating diameter from the expression [16]. The specific speed takes into account rotational speed, flow rate, and total mass-specific work extracted from the fluid, where 𝑁 is in units of rotations/second. For a water turbine with incompressible flow, the mass-specific work done on the fluid from leading to trailing edge, 𝑒, can be expressed in terms of total available head 𝐻. Specific speed 𝑁𝑠 is defined below in equation 1-60 [15]: 1 𝑁𝑠 = 1 𝜙′2 3 𝜓′4 = 𝑉̇ 2 ( 3) 𝑁𝐷 3 4 𝑒 ( ) (𝑁𝐷)2 1 = 𝑁𝑉̇ 2 3 𝑒4 1 = 𝑁𝑉̇ 2 3 𝑔𝐻 4 (1-60) It should be mentioned the head in this case is the total designed head of the turbine system, including losses. This form of specific speed is in units of rotations [4] . Another form of specific speed, Ω𝑠 , uses the angular speed of the turbine measured in radians per second instead of rotations per second [4], defined below for incompressable flow in equation 1-61: 29 1 Ω𝑠 = 𝜔𝑉̇ 2 (1-61) 3 (𝑔𝐻)4 Horlock stated there is an optimal value of the specific speed for a given type of machine, independent of size, at which efficiency is highest [15]. Figure 1.15 illustrates this concept, showing there is a specific speed at which peak efficiency is reached. Figure 1.15. Example efficiency curve for turbine for specific speed 𝑁𝑠 [15] 1.10.4. Specific Diameter Specific diameter, 𝐷𝑠 , is also used to describe and categorize turbines. The specific diameter was defined by using the head and capacity coefficients, similar to the specific speed. The head and capacity coefficients can be raised to a power to eliminate rotational speed from the equation [4]. This is shown below in equation 1-62 [4]: 1 𝐷𝑠 = 𝜓′4 1 𝜙′2 1 = 𝐷(𝑔𝐻)4 1 (1-62) 𝑉̇ 2 1.11. Cordier Diagram and Line Dixon stated that once specific speed is determined; the ideal machine type can be selected using the Cordier diagram [4]. The Cordier diagram is also often used to select a 30 specific speed or specific diameter for a given type of machine where efficiency is predicted to be highest. Otto Cordier carried out an experimental analysis of high efficiency turbomachines during the 1950s, and placed their data on a plot showing specific speed Ωs against specific diameter 𝐷𝑠 , and defined a trend line on a diagram showing where the highest efficiency turbomachines lie [17]. This curve is referred to as the “Cordier line”. Each type of machine has a range of specific speeds in which they perform with highest efficiency [15]. Machines with high head and low flow are on the right side of the Cordier diagram, while low head, high flow machines are on the left. The Cordier diagram with Cordier line is shown below in Figure 1.16. Circles are added to indicate where different types of machines typically lie on the diagram. Lines from Wright are placed above and below the Cordier line which indicate the accuracy/margin [18]. Wright divided the Cordier diagram into six regions [18]. The first region A, with specific speed between 6 and 10 and specific diameter between .95 and 1.25, propeller-type machines are typically used. Region B has specific speed between 3 and 6, specific diameter between 1.25 and 1.65. Wright mentioned this region has primarily axial turbomachines, like axial fans, axial pumps, and shrouded propellers. Region C contains machines with speed between 1.8 and 3, and specific diameter between 1.65 and 2.2. Ducted axial machines or multistage axial machines typically operate in this region. Region D includes machines with specific speeds between 1.0 and 1.8, and specific diameter between 2.2 and 2.8. Mixed flow pumps, blowers, mixed flow hydraulic turbines operate in this region. Region E includes machines with specific speed between .7 and 1, and specific diameter between 2.8 and 4. Centrifugal fans, pumps, heavy duty blowers, compressors operate in this range. Region F has specific speeds below .7 and specific diameters above 4. High pressure blowers, centrifugal compressors, high head pumps operate in this region. 31 Ωs Axials Mixed flow Radial D𝑠 Figure 1.16. 𝐷𝑠 − 𝛺𝑠 diagram with Cordier line and typical machine type Balje compiled test data and specifications from 92 water turbines built between 1940 and 1974 and placed them on the Ds − Ωs chart, superimposed over the Cordier line,. The water turbines included in that test data have specific speeds between 2 and 6. The water turbines are close to the line originally defined by Cordier [15]. The below Figure 1.17 shows the Cordier line alongside eight modern axial hydro turbines manufactured by Voith Hydro, the specifications of which were obtained on the company website [20]. 32 Specific Speed Ω𝑠 vs. Specific Diameter 14.00 12.00 Sp. Speed Ω𝑠 10.00 8.00 Voith Hydro 6.00 Cordier 4.00 2.00 0.00 0.0000 0.5000 1.0000 1.5000 2.0000 2.5000 3.0000 Specific Diameter Ds Figure 1.17. 𝐷𝑠 −𝛺𝑠 diagram with Voith Hydro turbines with Cordier line 1.11.1. Effect of Flow, Loading Coefficients on Cordier Diagram Position A turbine design’s position on the Ds −Ωs diagram relative to the Cordier line can be fully defined by two dimensionless coefficients, Flow coefficient 𝜙 and blade loading coefficient 𝜓. The below figures show the Cordier diagram with lines of constant 𝜙 (Figure 1.18) and 𝜓 (Figure 1.19) superimposed. 33 𝐷𝑠 vs. Ω𝑠 10 Cordier Sp. Speed Ω𝑠 FC = 0.005 FC = 0.01 FC=0.05 1 FC=0.1 1 10 FC = 0.25 FC=0.5 FC=1 FC=1.5 FC=3 0.1 Specific Diameter Figure 1.18. 𝐷𝑠 −𝛺𝑠 diagram with Cordier line, lines of constant flow coefficient 𝜙 𝐷𝑠 vs. Ω𝑠 10 Cordier LC=2 Sp. Speed Ω𝑠 LC=1 LC=0.5 LC=0.25 1 1 10 LC=0.1 LC=0.05 LC=3 0.1 Specific Diameter Ds Figure 1.19 𝐷𝑠 −𝛺𝑠 diagram with Cordier line, lines of constant blade loading coefficient 𝜓 Turbine designs with high blade loading and flow coefficients have both low specific speed and specific diameter, are located on the lower left side of the Ds −Ωs diagram, below and to the left of the Cordier line. Designs with low loading and flow coefficients are located above and to the right of the Cordier line. 34 1.11.2. Using Cordier Diagram as a Design Tool Wright recommended using the Cordier line as a design tool to help decide what type of machine to use for a particular application, or to pick optimal values of either specific diameter or specific speed given one of the two. In Fluid Machinery, Performance and analysis, Wright stated: “When it is necessary to meet certain requirements in volume flow rate and pressure rise, one can follow the overall trend prediction of the Cordier diagram to help decide what type of machine will perform the job. For an initial selection of speed and known fluid density, one can estimate the specific speed. Enter the diagram at Ω𝑠 , and determine a workable value for the machine size by choosing a value near or on the “Cordier line”. This initial procedure will indicate both the size and type of machine one should be considering and the highest value of efficiency one can reasonably expect to achieve”. [18] Wright stated that by using designs on the Cordier line highest efficiency can be achieved: “The efficiency band represents the best total efficiency that can reasonably be expected” [18]. Balje stated that efficient turbomachines are located close to the Cordier line [19]. Balje applied a curve fit to the data to derive an equation to pick optimal specific speed given specific diameter, which place the design on the Cordier line, the equations are shown below in equations 1-63 and 1-64 below [19]: 𝛺𝑠 ≅ 9.0𝐷𝑠−2.103 for 𝐷𝑠 < 2.8 (1-63) 𝛺𝑠 ≅ 3.25𝐷𝑠−1.126 for 𝐷𝑠 ≥ 2.8 (1-64) Adhikari et. al. described using the Cordier line as a starting point to help pick axial water turbine diameter given specific speed [21]. Given a flow rate, head, and desired diameter, the Cordier line can be used to pick an optimal specific speed, which then defines the RPM of the turbine. For example, for a turbine with a rotor diameter of .4 m, head of 2 meters, and flow rate of .46 𝑚3 𝑠 , the specific diameter is 1.22. Then the Cordier line recommendation is used to select 35 the specific speed of 5.9. This gives a rotational speed of 737 RPM. Doing this in reverse gives a recommended diameter given rotational speed, head, and volume flow rate. For example, a 170 RPM turbine with head of 2 meters, a volume flow rate of .46 𝑚3 𝑠 , specific speed is 1.37. Using the Cordier diagram to pick optimal specific diameter of 2.2. Using equation 1-62 and solving for diameter gives an outside diameter of .7 meters. 1.11.3. Operating below the Cordier Line The Cordier line can be used to pick a recommended specific diameter or specific speed given one of the two, however using the Cordier line for CTPAT turbines for this purpose results in recommendations of specific speed that may be larger than necessary for the application. Although there are correlations available within literature for predicting efficiency of specific types of hydraulic turbines at points off of the Cordier line, there will be no literature available for turbines with a new design, so it can be advantageous to test the effect of specific speed on efficiency when considering a new type of machine, like CTPAT. Although literature suggests that the Cordier line provides a reasonable estimation of the specific speed at which a given machine will perform at peak efficiency. Korpela noted that turbines with low specific speed are better suited for radial turbines, and as specific speed increases axial turbines perform with higher efficiency [16]. Dixon mentioned that the Cordier line is a mean curve based upon data from a large number of machines Cordier compiled, and “it is possible to diverge from the line and still obtain high performance pumps, fans and compressors” [4]. However Cordier and Balje’s recommendations suggest that turbomachines should perform with higher efficiency when they lie on the Cordier line. Therefore it makes sense to test whether this applies for CTPAT turbines. 36 This work investigates the viability of the CTPAT turbine scheme for low head hydro, at specific speeds above and below that recommended by the Cordier line. This work investigates whether designs of this type within the range of specifications tested below the Cordier line (lower than recommended specific speed) perform any worse than designs on the Cordier line. 1.12. Literature Recommendations on Blade Loading and Flow Coefficients Water turbines are often categorized by their blade loading coefficient 𝜓 and flow coefficient 𝜙. The Smith efficiency chart, shown below in Figure 1.20, was compiled using a large set of gas turbine test data to correlate head coefficient 𝜓′ and flow coefficient 𝜙 to efficiency for a turbines with constant axial velocity [22]. Each turbine was tested at different operating points to find its point of highest efficiency, and the flow coefficient and loading coefficient were plotted at that point. Although the original diagram was published in 1965, the Smith diagram is still used today for preliminary turbine design, as more modern turbine designs follow the same efficiency vs. flow and head coefficient trends [23]. Figure 1.20. Smith loading vs. Flow coefficient diagram [22] 37 Smith’s work suggests that as flow coefficient 𝜙 and head coefficient 𝜓′ are increased, efficiency tends to decrease. The same relationship applies for capacity coefficient 𝜙′ and the blade loading coefficient 𝜓. Turbines with higher flow coefficients have higher flow velocities relative to blade speed, this tends to increase losses as there is a smaller acceleration through the blade path, that acceleration is usually beneficial as it minimizes boundary layer growth or secondary flows [24]. When head coefficient is increased, the change in flow tangential velocity becomes larger, which tends to increase loss. [4] The Smith diagram is often used to choose optimum values of flow coefficient and blade loading coefficient given one of the two coefficients. Subsequent tests by Kacker and Okapuu confirmed the chart is useful for preliminary turbine design [25]. The diagram recommends values of flow coefficient and head coefficient that can offer higher performance, less flow separation and losses. However this does not mean that values of flow coefficient and head coefficient have to lie within recommended ranges in order to be considered acceptable designs [24]. The specifications of the axial turbines simulated lie both inside and outside of the recommendations of the Smith diagram and the Cordier line. Smith’s diagram was derived using gas turbine data and compressible flow, however the same trend applies for water turbines and incompressible flow, increasing flow coefficient and loading coefficient is correlated with decreased efficiency [15]. Figure 1.21 shows the range of typical flow coefficient 𝜙 and and blade loading coefficients 𝜓 for different types of hydraulic turbines from Horlock. Flow coefficients for Kaplan and Francis turbines rarely reach above .6, and blade loading coefficients rarely reach above 2 [15]. 38 Figure 1.21. Blade loading and flow coefficients for different types of hydraulic turbines [15] 1.13. Cost Advantages of Designs With High Flow and High Loading Coefficients Losses tend to increase with increasing blade loading and flow coefficients, however turbines with higher blade loading and flow coefficients can be produced for lower cost. Turbines with high flow coefficient generally have a smaller cross-sectional area compared to those with low flow coefficients, so turbines with lower 𝜙 tend to have increased turbine machining costs and weight [23]. Turbines with higher loading coefficient require fewer stages to achieve the same work extraction, and can extract more head for the same number of stages. Reducing the number of stages required for a turbine decreases part count, reducing weight, capital cost and operating cost. Industrial gas turbines are usually designed to maximize efficiency, so they generally have low flow and loading coefficients. Aerospace engines are usually designed for higher blade loading and high flow coefficient as the engine weight is of highest importance [23]. For aero engines, optimum value of 𝜓 is between 1.5 to 2.5, and 𝜑 39 ranges from 0.8 to 1.2. This contrasts with industrial gas turbines which typically have 𝜓 between 1.0 to 1.5, and 𝜑 from 0.4 to 0.8 [23]. Smith, in his paper defining his diagram, stated that in practice the designer should use the highest possible feasible values of blade loading and flow coefficient, and has to “strike a compromise between the conflicting requirements of efficiency versus weight and cost” [22]. A critical part of turbine selection is to weigh turbine efficiency with capital cost, and designs which are less costly to produce, but have lower efficiencies can be considered acceptable for many applications. This is especially true for low-head hydro applications, the revenue generated from power is often not enough to justify large capital expenses for the most efficient turbines. 1.14. Zweifel Lift Coefficient The Zweifel coefficient, also referred to as the Zweifel lift coefficient or Zweifel tangential force coefficient, 𝑍, is often used to find the optimal blade spacing where losses are 𝑠 predicted to be lowest. The blade solidity, , is the ratio of the blade spacing 𝑠 to the axial blade 𝑏 length 𝑏. The solidity and has a large effect on the friction and losses resulting from deceleration on the suction side (diffusion) in the blade row. The Zweifel coefficient recommendations in literature can be used to pick a value of 𝑠 𝑏 at which it is predicted that minimum friction losses and losses resulting from diffusion will occur [4]. If the spacing between blades is too small, the flow follows the blade with low deviation, but losses from friction are higher. If the spacing between blades is too high however, friction losses are low but the load distributed per blade increases, and losses resulting from diffusion increase [4]. Diffusion is undesirable in a turbine blade row, this is because the adverse pressure gradient, especially for large fluid deflections, makes boundary layer separation more probable [4]. 40 The losses in the turbine from leading to trailing edge can be expressed in terms of the profile loss coefficient, 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 . Profile losses are a function of losses as the result of friction and diffusion, such that 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 ≅ 𝜉𝑓𝑟𝑖𝑐𝑡𝑖𝑜𝑛 + 𝜉𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 . Schoberi stated that for every turbine design there is a value of 𝑠 𝑏 where overall profile loss coefficient 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 is minimum [14]. Figure 1.22 below shows profile loss coefficient at differing blade spacing/chord length ratios, as the result of diffusion and friction. Figure 1.22. Profile loss coefficient vs. s/b ratio resulting from separation and friction [14] Zweifel stated that “the ratio of an ’actual’ to an ‘ideal’ tangential blade loading has an approximately constant value for minimum losses” [26]. 𝑌 denotes the actual tangential force per blade, previously defined in equation 1-43, and 𝑌𝑖𝑑 is the ideal tangential force per blade. The Zweifel coefficient 𝑍 is defined using the below equation 1-65: 𝑌 𝑌𝑖𝑑 𝑍= (1-65) The tangential force imparted per blade, 𝑌, is equal to the area between the pressure and suction curves in Figure 1.13, and can be calculated using equation 1-43. 𝑌 can also be expressed using conservation of momentum, assuming axial velocity stays constant from the leading to trailing edge. Equation 1-66 below shows 𝑌 expressed using conservation of momentum, in 41 ̇ and the change in relative tangential velocities from terms of mass flow rate per blade ∆𝑚 trailing edge to leading edge, shown below in equation 1-66 [27]: ̇ (𝑊𝑢2 − 𝑊𝑢1 ) 𝑌 = ∆𝑚 (1-66) The ideal tangential force per blade 𝑌𝑖𝑑 is determined using an idealized pressure distribution from leading to trailing edge across a blade, multiplied by the area on which it acts. Zweifel defined this idealized pressure distribution using the assumption that the total pressure at the leading edge 𝑃𝑡1 acts over the entire pressure side of the blade, meaning that velocity is assumed to be zero [14]. For this idealized pressure difference used for derivation of the Zweifel coefficient, it is assumed that the pressure is equal to 𝑃2 on the entire suction side of the blade, and it is assumed the pressure is constant across the surface with no diffusion. The pressure distribution across the blade is thus assumed to be a rectangular shape [14]. 𝑌𝑖𝑑 is calculated by multiplying this pressure difference by the surface area on which this force acts, the product of the axial blade length and the blade height, 𝑏ℎ𝑏 , applied below in equation 1-67. 𝑌𝑖𝑑 = (𝑃𝑡1 − 𝑃2 )𝑏ℎ𝑏 (1-67) Figure 1.23 below shows an example pressure distribution across a blade with idealized pressure distribution for calculation of Zweifel coefficient. The Zweifel coefficient is a measure of how close the area of the real pressure distribution on the blades is to the area of this idealized pressure distribution. 42 Figure 1.23. Idealized pressure distribution with actual pressure distribution for Zweifel coefficient calculation For incompressible flow the idealized pressure difference 𝑃𝑡1 − 𝑃2 is equal to the dynamic pressure at the trailing edge, 𝜌𝑊22 2 [14]. Plugging into equation 1-67, the expression for idealized tangential force is seen below in equation 1-68: 𝑌𝑖𝑑 = 𝑏ℎ𝑏 𝜌𝑊22 2 (1-68) It is assumed that meridional velocity 𝑊𝑚 stays constant from leading edge to trailing edge, thus 𝑊𝑚1 = 𝑊𝑚2. By plugging in the definition of the mass flow rate through one blade path, ∆𝑚̇, equation 1-44, into the definition the actual tangential blade force 𝑌 (equation 1-66), 𝑌 can then be calculated, below in equation 1-69. 𝑌 = 𝜌ℎ𝑏 𝑠𝑊𝑚 (𝑊𝑢2 − 𝑊𝑢1 ) (1-69) Equations 1-68 and 1-69 can be plugged into the definition of 𝑍 (equation 1-65). Using ⃗⃗⃗ (equation 1-16), the definition of 𝑍 can be rearranged the definition of relative flow velocity 𝑊 to be expressed in terms of relative flow angles, seen below in equation 1-70 [27]: 43 𝑍= ̇ (𝑊𝑢2 − 𝑊𝑢1 ) 𝑌 ∆𝑚 𝑊𝑚2 𝑠(𝑊𝑢2 − 𝑊𝑢1 ) = = 1 2 (𝑃𝑡1 − 𝑃2 )𝑏ℎ𝑏 𝑌𝑖𝑑 2 𝑊2 𝑏 (1-70) The Zweifel coefficient is proportional to the ratio of the blade spacing to the axial blade 𝑠 length, 𝑏. Zweifel coefficient is similar to the tangential force coefficient 𝐶𝑦 , originally defined 𝑠 in equation 1-51, simplified in equation 1-71 below. Unlike Z, 𝐶𝑦 is not dependent upon 𝑏, and 𝐶𝑦 uses mean relative velocity to normalize the tangential force. A set of designs with the same relative flow velocity components can have different values of Z depending on blade spacing and axial blade length, however the value of 𝐶𝑦 will stay constant across each design. 𝐶𝑦 = 𝑌 1 2 2 𝜌𝑊𝑚𝑒𝑎𝑛 𝑠ℎ𝑏 = 𝑊𝑚2 (𝑊𝑢2 − 𝑊𝑢1 ) 1 2 2 𝑊𝑚𝑒𝑎𝑛 (1-71) The Zweifel coefficient recommendations from literature can be used to find the 𝑠 𝑠 recommended value of 𝑏. After calculating the ideal 𝑏, the ideal 𝑏 can then be calculated using a 𝑠 known value of 𝑠. Equation 1-70 is rearranged to solve for 𝑏 in equation 1-72 below: 𝑠 𝑍𝑊22 =. 𝑏 2𝑊𝑚2 (𝑊𝑢2 − 𝑊𝑢1 ) (1-72) The Zweifel coefficient depends on the number of blades, axial blade length, blade spacing, and the blade angles, which are all interconnected. Blades which are closer together have lower Z, and results in higher weight, increased cost, reduced blade loading per blade, and increased friction losses due to increased surface area. Blades which are farther apart have higher Z, lower cost, lower weight, more tangential force distributed per blade, but lower friction losses, due to the decreased blade surface area [22]. Farther apart blades, therefore higher loading per blade can increase losses resulting from diffusion and adverse pressure gradients, while offering the benefit of lower friction. As 44 Zweifel coefficient is raised further the risk of flow separation increases [28] [4] [23]. Literature suggests that for the designs with higher Zweifel coefficient, losses via friction from the blade surfaces, hub and shroud are reduced, however the losses resulting from diffusion and trailing edge losses are higher [24] [28]. 1.14.1. Recommended Values of Zweifel Coefficient Zweifel suggested that for turbine cascades with incompressible flow at low Mach numbers, minimum profile losses occur and maximum efficiency can be reached when 𝑍 = .8 [26]. However since Zweifel’s original paper, additional work has been done to determine optimal values of 𝑍, after Zweifels initial work Pfiel demonstrated the optimal value of the coefficient varies from 𝑍 =.75 to 1.15 depending on flow deflection [30]. The experimental data used to determine the optimal value of Zweifel coefficient is based on dated turbine blade profiles, with moderate loading [4]. E.Dick noted the recommended value of the Zweifel coefficient does not change depending on the Mach number, and works for both incompressible and compressible fluids. He mentions that the optimal value of Zweifel coefficient is between 1.0 and 1.2, with separation beginning with values between 1.4 and 1.6 [27]. Although there is no agreement on a Zweifel coefficient that results in the highest efficiency, the literature suggests that for a given turbine design, the peak efficiency will be measured at a Zweifel coefficient somewhere between .75 and 1.2. 1.14.2. Industry Trend - Higher Zweifel Coefficients A major trend in the turbine industry is the design of blades with ever increasing Zweifel coefficients, usually achieved by decreasing the number of blades. The desire to use turbine designs with higher Zweifel coefficients is primarily motivated by the desire to decrease total cost of ownership of turbine equipment. Blade profile design with high Zweifel coefficients is 45 an area of active research, motivated primarily by the need to reduce part count and engine weight in commercial gas turbines, especially jet engines. [31] Reducing part count reduces capital cost, and service cost, reducing total cost of ownership. Decreasing the number of blades increases the tangential force 𝑌 imparted per blade, this increases Zweifel coefficient. The definition of the blade spacing, equation 1-33, and equation 1-72 show that reducing the number of blades increases blade spacing 𝑠, in turn increasing Zweifel coefficient. Reducing the number of blades reduces the number of parts and lowers material costs, which in turn reduces manufacturing and maintenance costs [31]. This is considered important, especially in the aviation industry where engine weight is critical. Zweifel coefficient can also be increased by reducing the axial length of the blades, 𝑏. Reducing the axial blade length has a similar effect on the tangential loading to reducing the number of blades, this is because Zweifel coefficient is a measure of the tangential loading experienced per blade, weighed against the axial length. Reducing axial blade length is advantageous from a cost perspective as well, as blades with lower axial length require less material, cost less to manufacture, and weigh less. Tests involving the highest Zweifel coefficient blade profiles demonstrated (in a cascade) were conducted by Praisner et. al, who tested airfoils with Zweifel coefficients between 1.6 and 1.82 [32]. However there has been less success in demonstrating airfoils with very high Zweifel coefficients in actual commercial gas turbines [28]. Schmitz et. al. developed a turbine stage called the Notre Dame Highly Loaded Turbine 01 (ND-HiLT01), with a rotor Zweifel coefficient of 1.35, with a blade loading coefficient of 2.8, and measured a stage efficiency of 90.6% [31]. The ND-HiLT01 was developed to demonstrate a reduction in stage count for a gas turbine engine and the part count of an individual airfoil row. 46 1.14.3. Zweifel Analysis Reducing the number of blades and reducing the axial blade length both increase Zweifel coefficient for a design. The below Figure 1.24 shows how changing axial blade length 𝑏 affects Zweifel coefficient, for a turbine with same power and flow rate as the first simulation case in this work. The designs in the Blade Length Study in this work had their Zweifel coefficients adjusted in this way by changing 𝑏. b vs Z Zweifel Coefficient 2.5 2 1.5 1 0.5 0 0 0.05 0.1 0.15 0.2 Blade length b (m) Figure 1.24. Blade length vs. Zweifel coefficient Zweifel coefficient can also be changed by changing the number of blades of the design, which effects the blade spacing 𝑠. Reducing the number of blades increases the tangential force 𝑌 which is distributed per blade, thus increasing 𝑍. The below Figure 1.25 shows this effect for a turbine design. 47 Zweifel Coefficient Number of blades vs. Z 2.5 2 1.5 1 0.5 0 0 5 10 15 Number of blades Nb Figure 1.25. Number of blades vs. Zweifel coefficient Zweifel coefficient can be changed while keeping relative flow angles, mass flow rate, and power constant by changing the number of blades or the blade length. Zweifel coefficient can be changed by changing the tip speed 𝑈, however the relative flow angles will change. The tip speed 𝑈 has a significant effect on Zweifel coefficient, designs with larger tip speeds have lower Zweifel coefficients. The below Figure 1.26 shows how changing 𝑈 effects 𝑍 for different turbine designs with the same power output, with three different blade lengths. Zweifel Coefficient Tip Speed U vs Z 2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 b=0.06 b=0.02 b=0.1 0 5 10 15 20 Tip Speed U (m/s) Figure 1.26. Tip speed vs. Zweifel coefficient at differing blade lengths The above figure illustrates that designs with high tip speeds have low values of Zweifel coefficient. As tip speed is increased for a given design, to keep Zweifel coefficient near typical 48 levels between ..75 and 1.2, blade length has to be decreased. This was demonstrated in the simulations done in the Speed Study, in particular the designs with the tip speed recommended by the Cordier diagram. 1.15. Cavitation Cavitation occurs when the static pressure of the flow drops below the vapor pressure. Vapor bubbles are created by this effect, and these bubbles can collapse suddenly if the pressure rises later in the fluid stream. These collapsing bubbles can severely damage water turbines blades if the bubble collapse occurs on the turbine blade. The collapsing bubbles produce shock waves and microjets which produce high pressures and temperatures in a short period of time. Over time this effect causes fatigue in the blade material causing erosion [33]. Erosion resulting from cavitation occurs in pumps, water turbines, propellers, and valves. Cavitation is often generated on the suction surface of hydraulic turbine blades. 1.15.1. Cavitation for Composites Composites are considered an attractive material choice for turbine blades, as the high specific strength and stiffness of these materials are superior to that of metal blades [34]. The cost of composite materials is also becoming cheaper, leading turbomachinery manufacturers to move to composite blades in many applications. The Woven Wheel technology, which is constructed out of continuous composite fiber, could be employed in a hydraulic turbine application using a CTPAT setup, is useful to investigate literature to see how composites can perform under cavitation conditions. Studies by Yamatogi et al. investigated the cavitation resistance of composite propellers and the mechanism by which composite impellers are damaged by cavitation. Samples of three types of composites using reinforced fibers were tested under cavitation conditions, carbon fiber, 49 glass fiber, and aramid fibers. Specimens made of epoxy resin and aluminum bronze molded NAB (CAC703) were also tested. The study demonstrated the aramid fiber reinforced composite materials exhibited less erosion than carbon or glass fibers under cavitation. It was found the resistance to cavitation erosion was superior in situations where the adhesion between fiber and resin was stronger [34]. A study by M. Ćosić, M. Dojčinović & Z. Aćimović-Pavlović measured the cavitation resistance of aluminum matrix composite with silicon carbide reinforcement particles, and found the mass loss as the result of cavitation erosion was close to the mass lost for the same test performed for CA6NM stainless steel (a 13Cr–4Ni soft martensitic stainless steel). CA6NM is known for good corrosion and cavitation resistance, and is commonly used in hydraulic machinery [33]. These studies suggest that most composite blades are more susceptible to corrosion via cavitation than metal, and careful choice of fiber and resin is required to use composites for hydro propellers. 1.16. Discussion of Loss Mechanisms Figure 1.27. Trailing edge of blade with separated boundary layer [32] 50 Denton described the mechanisms surrounding viscous friction and entropy production in turbomachines [35]. Viscous friction is the result of viscous shear, which can occur in either boundary layers or mixing processes. A large portion of entropy generation in turbomachines is due to viscous shear, when the fluid undergoes a rate of shear strain. Velocity gradients which cause viscous shear are experienced at the boundary layers, the leading and trailing edges, and anywhere where flow separation is seen. Viscous shear stress is defined below in equation 1-73, where 𝑢 is the local flow velocity, 𝜏 is the local shear stress, and 𝑦 is position along a boundary layer [35]: 𝜏(𝑦) = 𝜇𝜕𝑢 𝜕𝑦 (1-73) The rate of entropy generation per unit surface volume in a boundary layer is shown below in equation 1-74, where 𝑇 is the local temperature [35]: 𝑆̇ = 1 𝑑𝑢 𝜏 𝑇 𝑑𝑦 (1-74) Entropy production in the boundary layer is proportional to the velocity cubed. This is why the entropy is generated more rapidly on the suction side of turbine blades than on the pressure side, where velocity is higher. Most boundary layers have velocity changing the most rapidly in the inner part of the boundary layer, especially for turbulent flow, the inner layer is where most entropy is generated within a boundary layer [32]. The wake left by the trailing edge of a turbine blade is considered a major source of losses. When two fluid streams at different velocity, pressure, or temperature mix together, entropy is generated. High shearing rates occur in wakes left after the trailing edge. Figure 1.27 above shows an example turbine blade trailing edge with separated boundary layer. Denton estimated the loss that can be attributed to the trailing edge can be estimated to be about 32% of 51 the boundary layer losses or 21% of the total losses in a subsonic turbine blade row. Thicker blades are correlated with additional losses at the trailing edge [35]. Due to the adverse pressure gradient arising from flow diffusion on the suction surface downstream of the minimum pressure, boundary layer separation near the trailing edge can occur, and is considered a major source of blade profile losses [35]. As flow moves from leading to trailing edge along the suction side and decelerates, the adverse pressure gradient can lead to backflow in the boundary layer. Flow separation can occur when there is backflow in the boundary layer, where local change velocity along the boundary layer 𝑑𝑢 𝑑𝑦 becomes negative. This can lead to a distinct wake region, and for blades with high Zweifel coefficient, flow separation is of added concern [29]. With higher Zweifel coefficients, pressure on the suction side is reduced and pressure on the pressure side is increased, the resulting increased difference in velocity from the suction side to the pressure side can lead to additional shear stress, entropy generation, and losses at the trailing edge. Turbine designs with higher Zweifel coefficients have a higher degree of flow deceleration on the suction side of the blades downstream the point of minimum pressure [24]. Another major source of loss is referred to as “secondary loss”, and includes friction losses resulting from the hub and shroud surfaces, as well as losses resulting from secondary flows near the hub and shroud [35]. Denton estimated that for turbines, the secondary losses are considered a major source of loss, contributing typically 1/3 of the overall loss [35]. 1.17. Goals of This Thesis and Description of Work This work explores the viability of using the CTPAT turbine design scheme for low-head hydroelectric applications, at speeds inside and outside of the recommendations of the Cordier line. This work also aims to evaluate the applicability of the literature recommendations of the 52 value of Zweifel coefficient for these types of machines, and to see how varying Zweifel coefficient by varying blade length effects performance. Operating with highest Zweifel coefficients requires careful design of the blade profile, as turbines with higher Zweifel loading coefficient are correlated with adverse pressure gradients on the suction side of the blades, and have an added risk of separation. In this work, the performance of the constant thickness blades in an axial turbine of the CTPAT type under varying Zweifel coefficients are tested using CFD. In the gas turbine industry, blade profiles are carefully designed using CFD in an effort to maximize performance at higher Zweifel coefficients. However detailed design of blade profiles can be computationally expensive and detailed precise blade profiles can be more expensive to machine. This work approaches this problem from a different perspective, using a simple blade profile and observing how it performs under differing conditions in simulation, with differing specific speeds and Zweifel coefficients. The first hypothesis tested is that designs of the type simulated with specific speed recommended by the Cordier line will perform with lower efficiency than designs with specific speed lower than that recommended by the Cordier line. This work investigated whether designs below the Cordier line (lower than recommended specific speed) perform with lower efficiency than designs of similar geometry, on the Cordier line. The second hypothesis tested is that peak efficiency will be obtained in CFD for designs with Zweifel coefficients in the range between 0.75 and 1.2, and decreased efficiency will be experienced at Zweifel coefficients above and below that range. To evaluate the applicability of the CTPAT design scheme to a low head hydro application and to evaluate these hypotheses, turbine designs were generated which defined the inputs and geometry for the CFD simulations. Losses resulting from a penstock pipe for a 53 hydroelectric dam retrofit application were estimated to estimate the total inlet pressure. Six sets of turbine specifications were investigated in CFD. First for the “Speed Study”, designs at each set of specifications at different rotational speeds were developed to test the performance of designs inside and outside of the Cordier recommendations. The highest efficiency designs from each case in the Speed Study were then simulated in the “Blade Length Study”, with varying axial blade lengths to investigate the effect of varying Zweifel coefficient on performance. The effect of varying specific speed and Zweifel coefficient on efficiency, trailing edge losses, friction losses from the blades and hub and shroud surfaces, flow deviation, and cavitation are investigated. Constraints were imposed on the geometry to ensure the designs fit the geometric and manufacturing constraints of the CTPAT design scheme and the Woven Wheel manufacturing process. The analytical model was set up to obtain CFD boundary conditions which were applicable for a proposed application of a CTPAT turbine for use as a retrofit unit to fit in front of dam spillways. Chapter 2 focuses on the analysis setup, the results of the CFD simulations are discussed in Chapter 3. 54 CHAPTER 2. ANALYSIS DESCRIPTION AND METHODS 2.1. Description of Setup An example application of a CTPAT turbine in a hydroelectric application is for hydroelectric dam retrofits, discussed in Chapter 1 of this work. The analytical model was set up to obtain CFD boundary conditions which were applicable for a proposed application of a CTPAT turbine for use as a retrofit turbine unit to fit on the front of dam spillways. Figure 2.1 below shows a simplified schematic of the setup. Turbine Figure 2.1. Hydroelectric dam retrofit schematic with height positions The heights defined by 𝑧a , 𝑧b , 𝑧1 , 𝑧2 , 𝑧c are input for this analysis. Each position is defined below.  Position a: water surface/inlet of penstock pipe  Position b: outlet of penstock pipe/inlet of turbine entry annulus  Position 1: Turbine leading edge/outlet of turbine entry annulus  Position 2: Turbine trailing edge /inlet of turbine exit annulus  Position c: Outlet of turbine exit annulus, water level 55 2.2. Analysis Description, Description of Specifications An analytical model was constructed in Microsoft Excel to define the inputs and boundary conditions needed for the CFD analysis. To simulate the designs in CFD, the below inputs are required. The purpose of the analytical model was to obtain these inputs.  Total pressure at inlet of the turbine entry annulus, 𝑃𝑡𝑏 , accounting for estimated friction losses in penstock pipe  Mass flow rate 𝑚̇  Tip speed of rotor 𝑈1 = 𝑈2  Relative blade angles at leading and trailing edges, 𝛽𝑏1 and 𝛽𝑏2 respectively  Density of water 𝜌  Geometry: o Inside and outside radius of turbine rotor 𝑟ℎ , 𝑟𝑠 respectively, equal to radii of entry and exit annuli o Blade thickness 𝑡 o Number of blades 𝑁𝑏 This model takes turbine and penstock geometry, head, and flow rates as inputs and calculates values of absolute and relative flow speeds/angles, estimated total pressure at the turbine entry annulus, as well as blade angles for CFD analysis. The analytical model was made to take account for losses in a penstock pipe, including friction resulting from the pipe surfaces, and the friction loss associated with a 90 degree turn of the penstock. Six different sets of specifications relevant to low-head hydro applications were investigated. Three size turbines were considered, at two different values of total available head, 𝐻. Each of the six sets of specifications had its own input volume flow rate. Designs at each set of specifications, referred 56 to as a “case”, were generated at differing rotational speeds for the Speed Study. After finding the RPM where efficiency was measured highest through CFD, a new set of designs were generated for each case, each with a different axial blade length and corresponding Zweifel coefficient. This second set of designs were simulated during the Blade Length Study. 𝑚3  Case 1: 𝐻 = 2𝑚, .2 m OD, .12m ID, 𝑉̇ = .46  Case 2: 𝐻 = 2𝑚, .3 m OD, .18 m ID, 𝑉̇ = 1.05  Case 3: 𝐻 = 2𝑚, .5 m OD, .3 m ID, 𝑉̇ = 2.85  Case 4: 𝐻 = 4𝑚, .2 m OD, .12m ID, 𝑉̇ = .6  Case 5: 𝐻 = 4𝑚, .3 m OD, .18 m ID, 𝑉̇ = 1.5  Case 6: 𝐻 = 4𝑚, .5 m OD, .3 m ID, 𝑉̇ = 4.1 𝑠 𝑚3 𝑠 𝑚3 𝑠 𝑚3 𝑠 𝑚3 𝑠 𝑚3 𝑠 The surface roughness of 314 stainless steel was used for the penstock pipe friction calculations. The temperature of the water flow was assumed to be 20 degrees Celsius for the calculations. This temperature was needed to find the kinematic viscosity and vapor pressure of the water flow. The analysis and CFD simulations did not include a draft tube, an important and necessary part of a turbine installation which can be used to decelerate the flow. For use in a real application a draft tube will need to be included as a part of the design. 2.3. Estimation of Friction Losses The purpose of this section is to demonstrate how the inputs for CFD described above were determined for the designs later simulated in CFD. For this analysis, Bernoulli’s equation was used to estimate the total work extraction from the fluid, 𝑒, which was needed to define the relative and absolute flow angles needed to generated designs for CFD analysis, as well as the total pressure at the inlet of the entry annulus, 𝑃𝑡𝑏 . Bernoulli’s equation is a modified version of 57 the energy conservation equation, and is defined below in equation 2-1, Where 𝑃 is static gage pressure, 𝐶 is absolute flow speed, 𝑧 is vertical distance from the datum, water level, 𝑔 is the gravitational acceleration constant, 𝐹𝑓 is the friction work done per unit mass of a fluid element while moving from positions 𝑖 and 𝑗 along a streamline in the direction of flow, and 𝑒 represents total mass-specific energy extraction from positions 𝑖 to 𝑗 [4]: 𝑃𝑗 𝐶𝑗2 𝑃𝑖 𝐶𝑖2 + + 𝑔𝑧𝑖 = + + 𝑔𝑧𝑗 + 𝐹𝑓 + 𝑒 𝜌 2 𝜌 2 (2-1) For the analysis in this work, 𝑒 was estimated by carrying out Bernoulli’s equation between each system position, a through c, then combining the equations to obtain an expression for 𝑒. This begins with points a and b, from the water surface to the outlet of the penstock pipe, shown below in equation 2-2.: 𝑃𝑎 𝐶𝑎2 𝑃𝑏 𝐶𝑏2 + + 𝑔𝑧𝑎 = + + 𝑔𝑧𝑏 + 𝐹𝑓𝑎𝑏 𝜌 2 𝜌 2 (2-2) For this analysis it is assumed that the velocity at the water surface is zero, and the gage static pressure at the water surface is zero. Thus 𝐶𝑎 , 𝑃𝑎 = 0. It is also assumed that the motion of the water at the outlet of the penstock pipe is purely axial in direction. The velocity of the water at the outlet of the penstock pipe can be calculated using equation 2-3 below, using the continuity equation. 𝐶𝑏 = 𝑚̇ 𝜌𝐴𝑚𝑎 (2-3) Mass-specific friction loss from positions a to b 𝐹𝑓𝑎𝑏 is calculated using the friction loss equation for flow in a pipe added to the friction loss equation due to a 90 degree bend, where 𝜉𝑝 is the friction loss coefficient for the penstock pipe, and 𝐾 is the friction loss coefficient resulting from the 90 degree turn, shown below in equation 2-4: 58 𝐹𝑓𝑎𝑏 = 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 + 2 2 (2-4) 𝜉𝑝 is found using the below equation 2-5, a re-ordered version of the Darcy-Weisbach equation, where 𝐿𝑡 is the length of the penstock pipe, 𝑓𝑎𝑏 is the friction factor for the penstock pipe, and 𝑟𝑏 is the radius of the penstock pipe [36]: 𝜉𝑝 = 𝑓𝑎𝑏 𝐿𝑡 2𝑟𝑏 𝑔 (2-5) There are many equations which can be used to estimate this friction factor for a fullflowing circular pipe, for this analysis the Haaland equation is used, valid for turbulent flow. The accuracy of the friction factor calculated using this equation is within ±2% for Reynolds numbers above 3000 [36]. First 𝑅𝑒𝑏 is calculated using equation 2-6 below, the Reynolds number at position b, where 𝜇 is the kinematic viscosity of water [4]: 𝑅𝑒𝑏 = 𝐶𝑏 𝑟𝑏 𝜇 (2-6) Then the Haaland equation is applied below in equation 2-7, 𝜀𝑟 is the surface roughness [36]. 𝑓𝑎𝑏 = 1 2 𝜀𝑟 1.11 6.9 2𝑟𝑏 (−1.8 log (𝑅𝑒 + ( 3.7 ) )) 𝑏 (2-7) The penstock pipe modeled as a part of this analysis turns 90 degrees, before the flow reaches the turbine leading edge. The losses resulting from turning the flow 90 degrees were predicted. The friction loss coefficient resulting from the 90 degree turn, 𝐾, is predicted using correlations from literature, specifically a chart for predicting 𝐾 for 90 degree bends of uniform diameter obtained from the Hydraulic Institute’s Pipe Friction Manual [37]. After obtaining the predicted friction factors, Equation 2-4 was then simplified and re-arranged to form equation 2-8 below: 59 𝑃𝑏 𝐶𝑏2 = 𝑔(𝑧𝑎 − 𝑧𝑏 ) − (1 + 𝜉𝑝 + 𝐾) 𝜌 2 (2-8) Then Bernoulli’s equation for positions b to 1 (from the inlet of the entry annulus to the leading edge of the turbine) is applied in equation 2-9 below. 𝑃𝑏 𝐶𝑏2 𝑃1 𝐶12 + + 𝑔𝑧𝑏 = + + 𝑔𝑧1 + 𝐹𝑓𝑏1 𝜌 2 𝜌 2 (2-9) The region between the inlet of the entry annulus to the leading edge of the turbine are simulated as a part of the CFD analysis. As such the friction loss from positions b to 1, 𝐹𝑓𝑏1 , is accounted for in the CFD simulation. The expression for 𝑃𝑏 𝜌 (equation 2-8) is then plugged into equation 2-9 to obtain equation 2-10 below: 𝑔(𝑧𝑎 − 𝑧𝑏 ) − 𝐶𝑏2 𝐶𝑏2 𝑃1 𝐶12 + 𝑔(𝑧𝑏 − 𝑧1 ) = + (1 + 𝜉𝑝 + 𝐾) + 2 2 𝜌 2 (2-10) Equation 2-10 is then simplified to obtain equation 2-11 below: 𝑔(𝑧𝑎 − 𝑧1 ) − 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 𝑃1 𝐶12 − = + 2 2 𝜌 2 Rearranging equation 2-11 to solve for 𝑃1 𝜌 (2-11) , equation 2-12 is formed: 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 𝐶12 𝑃1 = 𝑔(𝑧𝑎 − 𝑧1 ) − − − 𝜌 2 2 2 (2-12) Bernoulli’s equation is then applied for positions 1 to 2 (turbine leading edge to trailing edge) in equation 2-13 below. The total work extracted from the fluid by the turbine is 𝑒. 𝑃1 𝐶𝑐2 𝐶22 𝑃2 + + 𝑔𝑧1 = + + 𝑔𝑧2 − 𝑒 𝜌 2 2 𝜌 (2-13) Plugging equation 2-12 into equation 2-13 and rearranging, equation 2-14 below is formed, 𝑔(𝑧𝑎 − 𝑧1 ) − 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 𝐶22 𝑃2 − = + + 𝑔(𝑧2 − 𝑧1 ) + 𝑒 2 2 2 𝜌 60 (2-14) Bernoulli’s equation is then applied for positions 2 to c, from the turbine trailing edge to the outlet of the exit annulus, seen below in equation 2-15. 𝑃2 𝐶22 𝑃𝑐 𝐶𝑐2 + + 𝑔𝑧2 = + + 𝑔𝑧𝑐 + 𝐹𝑓2𝑐 𝜌 2 𝜌 2 (2-15) The region between the trailing edge of the turbine to the outlet of the exit annulus are simulated as a part of the CFD analysis. The friction loss from positions 2 to c, 𝐹𝑓𝑑𝑒 , is accounted for in the CFD simulations. As position c is located at the datum, water level, 𝑃𝑐 = 0. Rearranging equation 2-15 and simplifying, equation 2-16 below is formed: 𝑃2 𝐶𝑐2 𝐶22 = + 𝑔(𝑧𝑐 − 𝑧2 ) − 𝜌 2 2 (2-16) Plugging equation 2-16 into equation 2-14 and solving for 𝑒 gives equation 2-17 below: 𝑒=− 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 𝐶𝑐2 + 𝑔(𝑧𝑎 − 𝑧𝑐 ) − − 2 2 2 (2-17) With mass flow rate and geometry as inputs, absolute meridional flow speed at the leading edge 𝐶𝑚1 can be calculated using the continuity equation, applied below in equation 218: 𝐶𝑚1 = 𝑚̇ 𝐴𝑚 𝜌 (2-18) The turbine is designed so 𝐶𝑚 stays constant from leading to trailing edge, such that 𝐶𝑚1 = 𝐶𝑚2 . Flow angle at the leading edge 𝛼1 is also input, allowing 𝐶𝑢1 to be calculated using equation 1-19, applied below in equation 2-19: 𝐶𝑢1 = 𝐶𝑚1 tan 𝛼1 (2-19) Equation 2-17 is then rearranged to be expressed in terms of variables with known values. To simplify the estimation of 𝑒, 𝐶𝑐 , the velocity of the flow at the outlet of the exit annulus, is assumed to be equal to 𝐶2 , the velocity at the trailing edge. 𝐶2 is then expressed in 61 terms of the mass-specific turbine shaft work 𝑒̃ , by first expressing 𝐶2 in terms of its meridional and tangential velocity components, seen below in equation 2-20: 2 2 𝐶22 = 𝐶𝑚2 + 𝐶𝑢2 (2-20) Euler’s equation of turbomachinery, expressed in equation 1-40, is then applied between positions 1 and 2 (leading edge to trailing edge of turbine rotor), and then solved for 𝐶𝑢2 , shown below in equation 2-21: 𝐶𝑢2 = 𝑒̃ + 𝐶𝑢1 𝑈1 𝑈2 (2-21) 𝑒̃ Using 𝑒 = 𝜂, where 𝜂 is isentropic turbine efficiency, equation 2-21 is expressed in equation 2-22: 𝐶𝑢2 = 𝑒𝜂 + 𝐶𝑢1 𝑈1 𝑈2 (2-22) For the designs considered in this analysis, hub and shroud radii stayed constant from leading to trailing edge, such that 𝑈1 = 𝑈2 . For this analysis it was assumed that 𝐶𝑚 stays constant from leading to trailing edges, such that 𝐶𝑚1 = 𝐶𝑚2 . Equation 2-22 above is then plugged into equation 2-20 to express 𝐶2 in terms of known variables, shown below in equation 2-23: 𝐶22 = 2 𝐶𝑚1 𝑒𝜂 + 𝐶𝑢1 𝑈1 2 +( ) 𝑈2 (2-23) Equation 2-23 is then plugged into equation 2-17 to express 𝑒 in terms of known variables, shown below in equation 2-24: 2 𝐶𝑚1 +( 𝑒=− 𝑒𝜂 + 𝐶𝑢1 𝑈1 2 ) 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 𝑈2 + 𝑔(𝑧𝑎 − 𝑧𝑐 ) − − 2 2 2 62 (2-24) To determine the absolute and relative flow speeds and angles, a value of 𝜂 = 100% (ideal) was assumed, which allows for preliminary calculation of relative flow angles at the trailing edge that are necessary for generating designs for CFD. The quadratic equation is used to solve equation 2-24 for 𝑒. After 𝑒 was calculated, 𝐶𝑢2 needed to be determined in order to obtain the relative and absolute flow angles necessary for defining the turbine blade geometry. As the efficiency of the machine is not yet known, accurate values of 𝐶𝑢2 , 𝑒̃ , and 𝛽2 cannot yet be determined, and for this analysis are obtained via CFD simulations. The CFD simulations require inputting blade angles, 𝛽𝑏1 and 𝛽𝑏2 . For an ideal turbine, 𝑒̃ is equal to 𝑒, and 𝛽2 is then 𝛽𝑏2 = 𝛽2. However, the turbine will not perform at 100% efficiency, and the flow does not follow the blades perfectly, so the trailing edge relative flow angle 𝛽2 was reached in CFD by specifying an trailing edge blade angle 𝛽𝑏2 and adjusting it until the total mass-specific work extraction achieved through CFD was equal to 𝑒. An initial value of 𝛽𝑏2 was required as an input to conduct the first simulation, this is defined as 𝛽𝑏2𝑖 . Flow speeds and angles were calculated by first inputting an efficiency of 100% into equation 2-22 to obtain an initial, idealized value of absolute circumferential trailing edge velocity 𝐶𝑢2𝑖 . Then initial, idealized values of 𝑊2 , 𝑊𝑢2 , 𝐶2 , 𝛼2 , 𝛽1 were calculated, denoted in this work as 𝑊2𝑖 , 𝑊𝑢2𝑖 , 𝐶2𝑖 , 𝛼2𝑖 , 𝛽1𝑖 . Equations 1-18,1-31,1-17,1-19, and 1-20 were used, respectively, to calculate these initial values. Finally, equation 1-20 was used to obtain 𝛽𝑏2𝑖 , substituting 𝛽𝑏2𝑖 for 𝛽2 in the expression. An initial value of 𝛽𝑏1 = 𝛽1𝑖 was used for the first simulation. After conducting a first simulation with 𝛽𝑏2 = 𝛽𝑏2𝑖 , 𝛽𝑏2 was reduced iteratively until the desired value of 𝑒 was reached in CFD. After obtaining the desired work extraction through CFD, 𝑒̃ , 𝛽2, 𝐶𝑢2 , and efficiency 𝜂 were logged. 63 The calculations to obtain the flow speeds and angles were carried out at three positions along the blade: at the hub, shroud, and 50% span. Relative flow and blade angles were calculated such that the product of the rotor tangential speed multiplied by the ideal absolute tangential flow speed 𝑈𝜃 𝐶𝑢𝑖 is constant across the blade span at the trailing edge. The total pressure at the inlet of the turbine entry annulus, 𝑃𝑡𝑏 was then calculated, this value was used as a total pressure boundary condition for the annulus inlet for the CFD simulations. To simplify the analysis the gravitational potential energy of the turbine 𝜌𝑔(𝑧𝑐 − 𝑧𝑎 ) was applied in the form of total pressure at the annulus inlet for CFD simulations 𝑃𝑡𝑏 . This is shown below in equation 2-25, where 𝐹𝑓𝑎𝑐 is the sum of the friction losses considered above per unit mass from points a to c. 𝑃𝑡𝑏 = 𝜌𝑔(𝑧𝑎 − 𝑧𝑐 ) + 𝑃𝑎𝑡𝑚 − 𝑔𝐹𝑓𝑎𝑐 𝜉𝑝 𝐶𝑏2 𝐾𝐶𝑏2 = 𝜌𝑔(𝑧𝑎 − 𝑧𝑐 ) + 𝑃𝑎𝑡𝑚 − 𝜌 −𝜌 2 2 (2-25) 2.4. CFD Methodology and Assumptions 2.4.1. CFD Overview The analytical model was used to obtain the boundary conditions and inputs for computational fluid dynamics (CFD) software. The software package used was ANSYS. The CFD analysis was carried out in ANSYS CFX. Computational Fluid Dynamics (CFD) software solves modified forms of the Navier-Stokes equations, as well as continuity. The Navier-Stokes equations describe how the flow velocity, pressure, temperature, and density of a viscous fluid are related. The Navier-Stokes equations in the x, y, and z direction are listed below for incompressible flow with constant viscosity, where 𝑢 represents the local velocity, equation 2-26 is in the x direction, equation 2-27 in the y direction, and equation 2-28 is in the z direction: 𝜌( 𝜕𝑢𝑥 𝜕𝑢𝑥 𝜕𝑢𝑥 𝜕𝑢𝑥 𝜕𝑃 𝜕 2 𝑢𝑥 𝜕 2 𝑢𝑥 𝜕 2 𝑢𝑥 + 𝑢𝑥 + 𝑢𝑦 + 𝑢𝑧 )=− + 𝜌𝑔𝑥 + 𝜇( 2 + + ) 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑥 𝜕𝑦 2 𝜕𝑧 2 64 (2-26) 𝜕𝑢𝑦 𝜕𝑢𝑦 𝜕𝑢𝑦 𝜕𝑢𝑦 𝜕 2 𝑢𝑦 𝜕 2 𝑢𝑦 𝜕 2 𝑢𝑦 𝜕𝑃 𝜌( + 𝑢𝑥 + 𝑢𝑦 + 𝑢𝑧 )=− + 𝜌𝑔𝑦 + 𝜇( 2 + + ) 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑦 2 𝜕𝑧 2 𝜌( 𝜕𝑢𝑧 𝜕𝑢𝑧 𝜕𝑢𝑧 𝜕𝑢𝑧 𝜕𝑃 𝜕 2 𝑢𝑧 𝜕 2 𝑢𝑧 𝜕 2 𝑢𝑧 + 𝑢𝑥 + 𝑢𝑦 + 𝑢𝑧 )=− + 𝜌𝑔𝑧 + 𝜇( 2 + + ) 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑧 𝜕𝑥 𝜕𝑦 2 𝜕𝑧 2 (2-27) (2-28) The continuity equation is below in equation 2-29 for incompressible flow: 𝜕𝑢𝑥 𝜕𝑢𝑦 𝜕𝑢𝑧 + + =0 𝜕𝑥 𝜕𝑦 𝜕𝑧 (2-29) The four equations are all coupled. Time dependent solutions of the Navier-Sokes equations are too computationally expensive to be feasible for most situations. Reynoldsaveraging is used to simplify the process of solving the equations, in a way such that small scale turbulent fluctuations do not need to be simulated. The Reynolds averaged Navier-Stokes equations are time-averaged versions of the Navier-Stokes equations. For Reynolds averaging, the velocity and scalar variables of the Navier-Stokes equations are converted into mean (time averaged) and fluctuating components, shown below in equation 2-30, where 𝑢𝑗 are the mean velocity components (j=1,2,3) and 𝑢𝑗 ′ are the fluctuating velocity components: 𝑢𝑗 = 𝑢𝑗 + 𝑢𝑗 ′ (2-30) Pressure and the other scalar quantities in the Navier-Stokes equations are converted into mean and fluctuating components as well. The variables in this form are substituted into the Navier-Stokes continuity and momentum equations to derive the Reynolds averaged NavierStokes (RANS) equations, shown below in index notation in equation 2-32, where 𝛿𝑗𝑘 is the Kroneker delta, the definition of which is listed in equation 2-31 [36]. 0 𝑓𝑜𝑟 𝑗 ≠ 𝑘 𝛿𝑗𝑘 = { 1 𝑓𝑜𝑟 𝑗 = 𝑘 (2-31) 𝜕𝑢𝑗 𝜕𝑢𝑗 1 𝜕 𝜕𝑢𝑗 + 𝑢𝑘 = (−𝑃𝛿𝑗𝑘 + 𝜇 − 𝜌𝑢𝑗′ 𝑢𝑘′ ) 𝜕𝑡 𝜕𝑥𝑘 𝜌 𝜕𝑥𝑘 𝜕𝑥𝑘 (2-32) 65 ′ The last term represents the Reynolds stresses, 𝜏𝑗𝑘 , defined below in equation 2-33. ′ 𝜏𝑗𝑘 = −𝜌𝑢𝑗′ 𝑢𝑘′ (2-33) The nonlinearity of the Navier-Stokes equations has the result that the velocity fluctuations still appear in the RANS equations in the Reynold stress term. To obtain forms of the RANS equations which contain only the mean velocity and pressure, the Reynolds stress term needs to be modeled in terms of mean flow, removing references to the fluctuating component of the velocity. The Reynolds stresses are modeled using the Boussinesq hypothesis, which relates the Reynolds stresses to mean velocity gradients [38]. Joseph Boussinesq introduced the concept of eddy viscosity, and proposed relating the Reynolds stresses to the mean velocity gradients. 𝜇𝑡 is the turbulence eddy viscosity, 𝑘 is the turbulence kinetic energy, 1 defined as 𝑘 = 2 𝑢𝑗′ 𝑢𝑘′ . Equation 2-34 below describes this relationship [38]: 𝑢′𝑗 𝑢′𝑘 = 𝜇𝑡 ( 𝜕𝑢𝑗 𝜕𝑢𝑘 𝜕𝑢𝑗 2 + )𝛿 ) − (𝜌𝑘 + 𝜇𝑡 𝜕𝑥𝑘 𝜕𝑥𝑗 3 𝜕𝑥𝑗 𝑗𝑘 (2-34) The Boussinesq hypothesis was used for the 𝑘 − 𝜖 and 𝑘 − 𝜔 turbulence models used in the simulations. The 𝑘 − 𝜖 and 𝑘 − 𝜔 turbulence models use two additional transport equations to compute 𝜇𝑡 as a function of 𝑘, 𝜔 and 𝜖. 𝜖 is the turbulence dissipation rate and 𝜔 is the specific dissipation rate. The 𝑘 − 𝜖 turbulence model is considered most useful for free-shear flows with smaller pressure gradients. The 𝑘 − 𝜔 turbulence model aims to model near-wall flow features more accurately than the 𝑘 − 𝜖 turbulence model. For the CFD simulations conducted as a part of this work, the Shear Stress Transport (SST) RANS turbulence model was used. The SST turbulence model combines the 𝑘 − 𝜖 and 𝑘 − 𝜔 turbulence models, where 𝑘 − 𝜔 is used in the inner region of the boundary layer, and 𝑘 − 𝜖 is used in the free shear flow. 66 2.4.2. CFD Methodology and Assumptions For all CFD simulations in this study, a single axial turbine stage was simulated. Each design simulated included an annulus entry and exit section, each with an axial length of 0.3 meters. The blade profiles for each CFD run was of a constant thickness with rounded leading and trailing edges. Blade designs were generated using BladeGen. Figure 2.2 below shows example turbine geometry, and Figure 2.3 below shows the example turbine geometry in Bladegen. Figure 2.2. Design 1j5 generated in BladeGen Figure 2.3. Example turbine geometry with entry and exit annuli An outer shroud was included in the simulations, eliminating tip leakage from the simulation. Meshes were generated using ANSYS TurboGrid. All simulations were steady state, 67 and were run until convergence was reached. To track convergence the total pressure at the annulus outlet and turbine efficiency were monitored until their values stabilized. All CFD runs were tested using the boundary condition set Total Pressure Inlet/Mass Flow Rate Outlet. Total pressure at the annulus inlet was input, taking into account the friction losses from the penstock pipe. Figure 2.2 shows renderings of one of the designs generated in BladeGen. The turbines simulated ranged in power output from 3KW to 137KW, with heads ranging from 2 to 4 meters. A number of different flow characteristics and performance metrics were measured for each CFD run. All data was recorded using CFX-Post. For the simulations within each case, the total work extraction from the fluid (counting both shaft work and losses) was kept constant between each design. This was achieved by changing the trailing edge blade angles 𝛽𝑏2until the total pressure at the outlet of the exit annulus 𝑃𝑡𝑐 reaches the desired value, atmospheric pressure. This allows for the efficiency and performance of each turbine to be compared with each other on equal ground. Blade thickness was also kept constant within each case. CFD simulations are not perfect recreations of a physical system, some assumptions need to be made to simplify the modeling, to reduce computation time. Gravity was not modeled in the simulations, instead the pressure as the result of gravitational force was applied to the inlet of the turbine entry annulus. The designs tested in CFD do not include an inlet guide vane. The CFD simulations assume a uniform flow velocity across the inlet which was assumed to have no pre-swirl, such that 𝐶𝑢1 = 0, 𝛼1 = 90°. A single flow path with blade was modeled in the CFX simulations, the outputs were dependent upon an assumption of symmetry across each flow path. Only water was simulated in the blade path, cavitation was not modeled, although areas of low pressure are tracked using ANSYS CFD-POST. The simulations assumed an inlet water 68 temperature of 20°C. The simulations did not take account for flow disturbances and turbulence resulting from the penstock pipe bend, friction losses were taken into account to obtain a total inlet pressure for CFD, as described in section 2.2. When employed in a physical application, a flow-guiding nose cone at the inlet and outlet of the annulus sections will be required, something the simulations in this work do not include. These simulations did not take surface roughness of the blades and walls into account, and assume no-slip wall conditions. Efficiency was obtained from the CFD results, in the form of isentropic efficiency. 𝑊̇ is the shaft power output, and was calculated in CFD by first multiplying the forces imparted on the blades in x,y, and z axis directions by the absolute flow velocities in those directions, then summing up the components and multiplying that by the number of blades. This is expressed below in equation 2-35: 𝑁𝑏 𝑊̇ = ∑(𝐹𝑥 𝐶𝑥 + 𝐹𝑦 𝐶𝑦 + 𝐹𝑧 𝐶𝑧 ) (2-35) 1 Isentropic efficiency was calculated in CFD-Post using the below equation 2-36, where 𝐻 is the total available head. 𝜂= 𝑊̇ 𝑉̇ 𝐻𝑔𝜌 (2-36) 2.5. Description and Inputs for Speed Study In this work, each set of specifications described in section 2.2 were referred to as a “Case”. Designs were generated and simulated at least 6 different rotational speeds for each case, in addition to the rotational speed recommended by the Cordier diagram. Idealized relative flow angles at the leading and trailing edges were calculated and used to define the initial blade shape. First a value of Z = .8 was used to determine the axial blade length at each rotational speed 69 tested. Each RPM had its own corresponding axial blade length 𝑏 which was adjusted to keep Zweifel coefficient at .8. In this analysis, preliminary geometries were first generated in the analytical model, with leading edge and trailing edge flow beta angles specified. For each rotational speed tested, new leading edge and trailing edge relative flow angles were calculated using the analytical model. For the first simulations, the leading edge blade angle 𝛽𝑏2 was kept the same as the ideal flow beta angle 𝛽2𝑖 . The total pressure at the annulus outlet will not reach the desired value, so for each design the trailing edge blade beta angle was reduced until the total pressure at the annulus outlet 𝑃𝑡𝑐 reached within 200 pA of the desired value, atmospheric pressure. When the total pressure at the annulus outlet reached the desired value, the data was recorded. For the Speed Study, Zweifel coefficient was kept at the original recommendation by Zweifel, Z = 0.8. This was achieved by using equation 1-72 to calculate axial blade length 𝑏, using a value of 0.8 for Z. Each of the six cases had a rotational speed at which the highest efficiency and power output was measured. The cases at which these geometries produced the highest power/efficiency were considered for the next analysis, the Blade Length Study. A flow chart describing the CFD simulation process for the Speed Study is shown in Figure 2.4 below. The below process was repeated for all six cases. 70 Figure 2.4. Process flow chart for Speed Study Table 2.1 below shows the geometric parameters, idealized flow speeds, flow angles, and ideal absolute circumferential flow speed at the trailing edge 𝐶𝑢2𝑖 which was used to obtain 𝛽2𝑖 for the first simulation, for two designs within Case 1 of the Speed Study, designs 1j and 1CORD. The table also shows the final values of 𝛽𝑏2 that were used to obtain the results. For the simulations of each design, 𝛽𝑏1 was kept equal to 𝛽1𝑖 , as first simulations showed incidence 𝑖 ′ was close to zero for every design simulated. Table 2.1 below shows how the higher speed designs in the Speed Study, in this case 1CORD, require a smaller axial blade length to keep Z at 71 .8. The designs were set up such that the product of the ideal circumferential flow speed and the circumferential rotor velocity, 𝑈𝜃 𝐶𝑢𝑖 was kept constant at all positions across the blade span. 𝑡 𝑟𝜃 𝑚 0.005 0.005 0.005 0.005 𝑚 0.200 0.160 0.165 0.120 𝑡 𝑟𝜃 𝑚 0.005 0.005 0.005 0.005 𝑚 0.200 0.160 0.165 0.120 Table 2.1. Case 1 Geometric Parameters and Flow Angles 1j 𝛽1𝑖 𝛽1𝑏 𝛽2𝑖 𝑈𝜃 𝐶𝑢1 𝐶𝑢2𝑖 𝑈𝜃 𝐶𝑢𝑖 𝛼2𝑖 𝛽𝑏2 2 𝑚 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑚/𝑠 𝑚/𝑠 𝑚/𝑠 𝑠2 3.6 0.0 -5.1 -18.0 129.9 59.1 59.1 34.6 26.8 2.8 0.0 -6.3 -18.0 136.1 64.4 64.4 33.0 25.2 2.9 0.0 -6.1 -18.0 135.3 63.8 63.8 33.3 25.5 2.1 0.0 -8.4 -18.0 144.0 70.3 70.3 29.6 21.8 1cord 𝑈𝜃 𝐶𝑢1 𝐶𝑢2𝑖 𝑈𝜃 𝐶𝑢𝑖 𝛼2𝑖 𝛽1𝑖 𝛽1𝑏 𝛽2𝑖 𝛽𝑏2 𝑚2 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑑𝑒𝑔 𝑚/𝑠 𝑚/𝑠 𝑚/𝑠 𝑠2 15.4 0.0 -1.2 -18.1 101.0 21.1 21.1 19.7 14.4 12.3 0.0 -1.5 -18.1 103.6 25.8 25.8 23.3 18.0 12.7 0.0 -1.4 -18.1 103.2 25.1 25.1 22.8 17.5 9.3 0.0 -2.0 -18.1 107.9 32.7 32.7 28.0 22.7 𝑏 𝑚 0.098 0.098 0.098 0.098 𝑏 𝑚 0.022 0.022 0.022 0.022 Z 0.8 0.8 0.8 0.8 Z 0.8 0.8 0.8 0.8 Velocity triangles of the designs were produced, Figures 2.5 and 2.6 show velocity triangles at leading and trailing edges for design 1j 1Cord of the Speed Study respectively, superimposed over the blade profile design produced using BladeGen. The higher speed designs, including 1Cord, have lower relative flow angles at leading and trailing edges, this is due to the higher rotational speed of the rotor for these designs. 72 Figure 2.5. Velocity triangles for leading and trailing edges of design 1j of the Speed Study (50% span) 73 Figure 2.6. Velocity triangles for leading and trailing edges of design 1CORD of the Speed Study (50% span) The below tables show the rotational speeds, axial blade length, and non-dimensional coefficients of the designs tested in CFD for each case for the Speed Study. Case 1 𝑚3 𝑉̇ = 0.46 𝑠 𝑟ℎ = .12 m 𝑟𝑠 = .2 m Table 2.2. Case 1 Speed Study Inputs Run Total Head (m) 𝑏 (m) RPM 𝜙 Ω𝑠 b 1.88 0.045 350 0.79 2.80 d 1.86 0.056 300 0.92 2.41 f h j l 1.86 1.84 1.85 1.88 0.070 0.086 0.098 0.106 250 200 170 140 74 1.12 1.39 1.63 1.98 2.01 1.62 1.37 1.12 𝐷𝑠 1.22 1.22 1.22 1.22 1.22 1.22 𝜓 0.34 0.46 0.66 1.03 1.43 2.14 Case 2 𝑚3 𝑉̇ = 1.050 𝑠 𝑟ℎ = .2 m 𝑟𝑠 = .3 m Case 3 𝑚3 𝑉̇ = 2.85 𝑠 𝑟ℎ = .3 m 𝑟𝑠 = .5 m Case 4 𝑚3 𝑉̇ = 0.6 𝑠 𝑟ℎ = .12 m 𝑟𝑠 = .2 m Case 5 𝑚3 𝑉̇ = 1.5 𝑠 𝑟ℎ = .2 m 𝑟𝑠 = .3 m Table 2.3. Case 2 Speed Study Inputs Run Total Head (m) 𝑏 (m) RPM 𝜙 Ω𝑠 b 1.90 0.031 350 0.53 4.19 d 1.85 0.044 300 0.62 3.67 f 1.88 0.060 250 0.75 3.01 h 1.88 0.086 200 0.94 2.42 j 1.88 0.105 170 1.10 2.05 l 1.85 0.126 140 1.34 1.71 n 1.87 0.142 120 1.56 1.45 Table 2.4. Case 3 Speed Study Inputs Run Total Head (m) 𝑏 (m) RPM 𝜙 Ω𝑠 f 1.90 0.034 250 0.45 4.92 h 1.89 0.061 200 0.56 3.96 j 1.92 0.104 150 0.74 2.94 l 1.91 0.148 120 0.93 2.36 n 1.88 0.199 90 1.24 1.79 p 1.88 0.260 60 1.86 1.19 Table 2.5. Case 4 Speed Study Inputs Run Total Head (m) 𝑏 (m) RPM 𝜙 Ω𝑠 b 3.73 0.05 450 0.80 2.45 d 3.72 0.06 400 0.90 2.19 f 3.68 0.07 350 1.03 1.93 h 3.70 0.08 300 1.20 1.64 j 3.70 0.09 250 1.45 1.37 l 3.70 0.10 220 1.64 1.21 Table 2.6. Case 5 Speed Study Inputs Run Total Head (m) 𝑏 (m) RPM 𝜙 Ω𝑠 b 3.73 0.04 450 0.59 3.88 d 3.67 0.05 400 0.67 3.49 f 3.71 0.06 350 0.76 3.03 h 3.70 0.08 300 0.89 2.60 j 3.72 0.10 250 1.07 2.16 l 3.72 0.11 220 1.21 1.90 n 3.70 0.13 190 1.41 1.65 p 3.72 0.15 160 1.67 1.38 75 𝐷𝑠 1.22 1.21 1.21 1.21 1.21 1.21 1.21 𝜓 𝐷𝑠 1.23 1.23 1.23 1.23 1.23 1.23 𝜓 𝐷𝑠 1.27 1.27 1.27 1.27 1.27 1.27 𝜓 𝐷𝑠 1.20 1.20 1.20 1.20 1.20 1.20 1.20 1.20 𝜓 0.15 0.20 0.30 0.47 0.65 0.94 1.29 0.11 0.17 0.31 0.47 0.83 1.87 0.41 0.52 0.67 0.92 1.33 1.71 0.18 0.23 0.30 0.41 0.59 0.76 1.02 1.44 Case 6 𝑚3 𝑉̇ = 4.1 𝑠 𝑟ℎ = .3 m 𝑟𝑠 = .5 m Table 2.7. Case 6 Speed Study Inputs Run Total Head (m) 𝑏 (m) RPM 𝜙 Ω𝑠 b 3.79 0.11 200 0.79 2.82 d 3.79 0.14 170 0.93 2.39 f 3.78 0.18 140 1.13 1.98 h 3.78 0.22 110 1.43 1.55 j 3.79 0.26 80 1.97 1.13 l 3.75 0.08 250 0.63 3.55 𝐷𝑠 1.22 1.22 1.22 1.22 1.22 1.22 𝜓 0.34 0.47 0.69 1.12 2.12 0.21 2.6. Description and Inputs for Blade Length Study The Blade Length Study was conducted to test the effect of changing Zweifel coefficient on turbine performance. For each turbine geometry and specified rotational speed, the axial blade length 𝑏 was systematically changed to alter the Zweifel coefficient. Designs with both short blades (large Zweifel coefficient) and long blades (small Zweifel coefficient) were simulated. Similar to the Blade Length Study, as blade length was changed 𝛽𝑏2 was changed and simulations were run until 𝑃𝑡𝑐 was within 200 Pa of 𝑃𝑎𝑡𝑚 . Figure 2.7 below is a flow chart which describes the steps required to obtain the desired data in CFD. 76 Figure 2.7. Blade Length Study process flow chart Velocity triangles of the designs were produced, Figure 2.8 and 2.9 shows velocity triangles at leading and trailing edges for designs 2l and 2l8 of the Speed Study respectively, superimposed over the blade profile design produced using BladeGen. 77 2l Z = 0.42 𝜂 = 82.9% Figure 2.8. Velocity triangles for leading and trailing edges of design 2l of the Blade Length Study (50% span) 78 2l8 Z = 0.84 𝜂 = 85.5% Figure 2.9. Velocity triangles for leading and trailing edges of design 2l8 of the Blade Length Study (50% span) At least eight designs with differing blade lengths were simulated for each case. Designs with blade lengths corresponding to Zweifel coefficients from .36 to 1.49 were simulated for the Blade Length Study. The below Table 2.7 shows each simulation which was run for the Blade Length Study, showing the blade length and Zweifel coefficient for each design. 79 Case 1j 1j 1j 1j 1j 1j 1j 1j Case 4h 4h 4h 4h 4h 4h 4h 4h 4h - Run 1 2 3 4 5 6 7 8 Run 1 2 3 4 5 6 7 8 9 - 𝒃 (𝒎) 0.162 0.122 0.090 0.070 0.050 0.190 0.080 0.096 𝒃 (𝒎) 0.16 0.14 0.18 0.11 0.08 0.06 0.09 0.05 0.07 - Table 2.8. Blade Length Study Runs and Inputs Z Case Run 𝒃 (𝒎) Z Case Run 0.48 2l 1 0.24 0.42 3n 1 0.63 2l 2 0.19 0.53 3n 2 0.85 2l 3 0.15 0.67 3n 3 1.09 2l 4 0.11 0.91 3n 4 1.49 2l 5 0.08 1.24 3n 5 0.41 2l 6 0.10 1.05 3n 6 0.96 2l 7 0.13 0.78 3n 7 0.80 2l 8 0.12 0.84 3n 8 9 3n Z Case Run 𝒃 (𝒎) Z Case Run 0.40 5n 1 0.24 0.44 1 6h 0.46 5n 2 0.27 0.39 2 6h 0.36 5n 3 0.20 0.52 3 6h 0.58 5n 4 0.16 0.65 4 6h 0.80 5n 5 0.12 0.87 5 6h 1.06 5n 6 0.10 1.03 6 6h 0.71 5n 7 0.08 1.28 7 6h 1.26 5n 8 0.11 0.94 8 6h 0.91 5n 9 0.11 0.80 9 6h 10 6h 11 6h 12 6h 𝒃 (𝒎) 0.40 0.35 0.30 0.25 0.20 0.15 0.13 0.11 0.18 𝒃 (𝒎) 0.40 0.44 0.47 0.50 0.37 0.33 0.29 0.25 0.21 0.16 0.13 0.20 Z 0.40 0.46 0.53 0.64 0.79 1.05 1.20 1.42 0.90 Z 0.44 0.41 0.38 0.36 0.48 0.54 0.61 0.70 0.83 1.08 1.31 0.90 2.7. Mesh Independence Study A mesh independence study was conducted first to ensure a fine enough mesh was used for each simulation to capture the flow features necessary to resolve the power, efficiency, and pressure difference accurately. Simulations with a finer mesh can resolve more flow details, but are more difficult to obtain convergence. A design was simulated for each case from the Speed Study, logging the performance characteristics, then the mesh size factor in ANSYS TurboGrid was changed systematically to see how mesh count affected power output. A base mesh count was picked at a point where increasing the mesh count further does not have a significant effect on power output. As each design has its own axial length, to ensure similarity across each design 80 simulated, mesh density was kept constant across each design, such that the cell count per meter stays constant across each design per case. Figure 2.10 below shows the power output for a single design measured with different cell counts in case 1, Figure 2.11 shows this for case 2, Figure 2.12 shows this for case 3, Figure 2.13 shows this for case 4, Figure 2.14 shows this for Power case 5, Figure 2.15 shows this for case 6. 7000 6950 6900 6850 6800 6750 6700 6650 6600 0 200000 400000 600000 800000 1000000 1200000 Cell Count Figure 2.10. Mesh cell count vs. Power for case 1 mesh Study Base mesh count of 550000 was used for the case 1 simulations, corresponding to an axial blade length of .14 meters. For case 1, simulations used a cell count of 3.9x10^6 cells/m. 17800 Power 17600 17400 17200 17000 16800 0 500000 1000000 1500000 Cell Count Figure 2.11. Mesh cell count vs. Power for case 2 mesh Study A base mesh count of 800000 was used for the case 2 simulations, corresponding to an axial blade length of .16 meters. For case 2, simulations used a cell count of 5.0x10^6 cells/m. 81 47500 Power 47000 46500 46000 45500 45000 44500 0 500000 1000000 1500000 Cell Count Figure 2.12. Mesh cell count vs. Power for case 3 mesh Study A base mesh count of 1000000 was used for the case 3 simulations, corresponding to an axial blade length of .3 meters. For case 3, simulations used a cell count of 3.3x10^6 cells/m. 18200 Power 18000 17800 17600 17400 17200 17000 0 500000 1000000 1500000 2000000 2500000 3000000 Cell Count Figure 2.13. Mesh cell count vs. Power for case 4 mesh Study A base mesh count of 1000000 was used for the case 4 simulations, corresponding to an axial blade length of .18 meters. For case 4, simulations used a cell count of 5.5x10^6 cells/m. 34000 Power 33500 33000 32500 32000 31500 0 500000 1000000 1500000 Cell Count Figure 2.14. Mesh cell count vs. Power for case 5 mesh Study 82 A base mesh count of 600000 was used for the case 5 simulations, corresponding to an axial blade length of .2 meters. For case 5, simulations used a cell count of 3x10^6 cells/m. 24500 Power 24000 23500 23000 22500 22000 0 200000 400000 600000 800000 Cell Count Figure 2.15. Mesh cell count vs. Power for case 6 mesh Study A base mesh count of 600000 was used for the case 6 simulations, corresponding to an axial blade length of .25 meters. For case 6, simulations used a cell count of 2.4x10^6 cells/m. 83 CHAPTER 3. CFD RESULTS AND DISCUSSION 3.1. Speed Study CFD Results The Speed Study tested designs with similar geometry at varying rotational speeds. Axial blade length was changed for each design such that 𝑍 = .8 for each design tested in the Speed Study. The below tables show how varying rotational speed effected efficiency for the different designs. Case 1 𝑉̇ = 𝑚3 𝑠 0.46 𝑟ℎ = .12 m 𝑟𝑠 = .2 m Table 3.1. Case 1 Speed Study CFD Results Efficiency 𝑈 𝑚2 𝑒̃ ( 𝑠2 ) 𝜂 𝐻(m) 𝑏(m) RPM (m/s) 1.88 0.045 14.38 77.09 350 7.33 1.86 0.056 14.60 79.04 300 6.28 Case Run 1 b 1 d 1 1 1 1 1 f h j l CORD. 1.86 1.84 1.85 1.88 2.07 0.070 0.086 0.098 0.106 0.022 14.69 14.69 14.77 14.45 6.45 79.66 80.47 80.58 77.53 34.87 250 200 170 140 737 5.24 4.19 3.56 2.93 15.44 𝜙 0.79 0.92 Ω𝑠 2.80 2.41 𝑊̇ (kW) 𝜓 0.27 6.62 0.37 6.72 1.12 1.39 1.63 1.98 0.38 2.01 1.62 1.37 1.12 5.47 0.54 0.84 1.17 1.68 0.03 6.76 6.76 6.79 6.65 2.97 Peak efficiency was observed at a specific speed of 1.4, RPM 170. Decreased efficiency was measured at rotational speeds above 300 RPM. Case 2 𝑚3 𝑉̇ = 1.050 𝑠 Case 𝑟ℎ = .2 m 2 𝑟𝑠 = .3 m 2 2 2 2 2 2 2 Run b d f h j l n CORD. Table 3.2. Case 2 Speed Study CFD Results 𝑒̃ Efficiency 𝑈 𝑚2 ( ) 𝐻(m) 𝑏(m) RPM (m/s) 𝜂 𝑠2 1.90 0.031 13.62 73.35 350 11.00 1.85 0.044 14.64 78.62 300 9.42 1.88 0.060 15.31 82.48 250 7.85 1.88 0.086 15.93 84.95 200 6.28 1.88 0.105 16.02 85.38 170 5.34 1.85 0.126 15.85 85.49 140 4.40 1.87 0.142 15.80 84.65 120 3.77 1.92 0.027 10.07 52.57 497 15.61 𝜙 0.53 0.62 0.75 0.94 1.10 1.34 1.56 0.38 Ω𝑠 4.19 3.67 3.01 2.42 2.05 1.71 1.45 5.91 𝜓 0.11 0.16 0.25 0.40 0.56 0.82 1.11 0.04 𝑊̇ (kW) 14.30 15.37 16.08 16.73 16.83 16.64 16.59 10.57 Peak efficiency was observed at a specific speed of 1.7, RPM of 140. A sharp decrease in efficiency was measured at rotational speeds above 250 RPM. 84 Table 3.3. Case 3 Speed Study CFD Results Case 3 𝑚3 𝑉̇ = 2.85 Case 3 3 3 3 3 3 3 𝑠 𝑟ℎ = .3 m 𝑟𝑠 = .5 m Run f h j l n p CORD. 𝐻(m) 1.90 1.89 1.92 1.91 1.88 1.88 1.84 𝑚2 𝑏(m) 𝑒̃ ( 𝑠2 ) Efficiency 𝜂 RPM 𝑈 (m/s) 𝜙 Ω𝑠 0.034 15.45 82.56 250 13.09 0.45 4.92 0.061 15.58 83.98 200 10.47 0.56 3.96 0.104 16.65 88.24 150 7.85 0.74 2.94 0.148 16.83 89.59 120 6.28 0.93 2.36 0.199 16.59 89.76 90 4.71 1.24 1.79 0.260 16.14 87.39 60 3.14 1.86 1.19 0.042 12.40 68.15 295 15.45 0.38 5.95 𝜓 0.09 0.14 0.27 0.43 0.75 1.64 0.05 𝑊̇ (kW) 44.04 44.41 47.44 47.97 47.28 46.01 35.35 Peak efficiency was observed at a specific speed of 1.8, corresponding to an RPM of 90. A sharp decrease in efficiency was measured at rotational speeds above 150 RPM. Case 4 𝑚3 𝑉̇ = 0.6 Case 4 4 4 4 4 4 4 4 𝑠 𝑟ℎ = .12 m 𝑟𝑠 = .2 m Run 𝐻(m) b 3.73 d 3.72 f 3.68 h 3.70 j 3.70 l 3.70 n 3.70 CORD. 3.73 Table 3.4. Case 4 Speed Study CFD Results Efficiency 𝑚2 𝑏(m) 𝑒̃ ( 𝑠2 ) RPM 𝑈 (m/s) 𝜂 0.05 29.43 79.62 450 9.42 0.06 29.89 81.05 400 8.38 0.07 29.92 81.96 350 7.33 0.08 30.11 81.99 300 6.28 0.09 29.81 81.22 250 5.24 0.10 29.30 79.98 220 4.61 0.05 28.61 77.94 500 10.47 0.03 13.95 37.98 997 20.88 𝑊̇ (kW) 𝜙 0.80 0.90 1.03 1.20 1.45 1.64 0.72 0.36 Ω𝑠 2.45 2.19 1.93 1.64 1.37 1.21 2.74 5.44 𝜓 0.33 0.43 0.56 0.76 1.09 1.38 0.26 0.03 17.66 17.93 17.95 18.07 17.89 17.58 17.16 8.37 Peak efficiency was observed at a specific speed of 1.6, RPM of 300. Decreased efficiency was measured at rotational speeds above 400 RPM. Case 5 𝑚3 𝑉̇ = 1.5 𝑠 𝑟ℎ = .2 m 𝑟𝑠 = .3 m Case 5 5 5 5 5 5 5 5 5 Run 𝐻(m) b 3.73 d 3.67 f 3.71 h 3.70 j 3.72 l 3.72 n 3.70 p 3.72 CORD. 3.68 Table 3.5. Case 5 Speed Study CFD Results Efficiency 𝑈 𝑚2 𝑏(m) 𝑒̃ ( 𝑠2 ) RPM (m/s) 𝜂 0.04 28.62 77.54 450 14.14 0.05 29.36 80.70 400 12.57 0.06 30.01 82.43 350 11.00 0.08 31.18 85.21 300 9.42 0.10 31.69 86.17 250 7.85 0.11 31.62 86.29 220 6.91 0.13 31.64 85.96 190 5.97 0.15 30.15 82.82 160 5.03 0.04 19.13 52.44 704 22.12 85 𝑊̇ (kW) 𝜙 0.59 0.67 0.76 0.89 1.07 1.21 1.41 1.67 0.38 Ω𝑠 3.88 3.49 3.03 2.60 2.16 1.90 1.65 1.38 6.13 𝜓 0.14 0.19 0.25 0.35 0.51 0.66 0.89 1.19 0.04 42.93 44.04 45.01 46.77 47.54 47.43 47.47 45.23 28.69 Peak efficiency was observed at a specific speed of 1.7, RPM of 190. A sharp decrease in efficiency was measured at rotational speeds above 300 RPM. Case 6 𝑚3 𝑉̇ = 4.1 𝑠 𝑟ℎ = .3 m 𝑟𝑠 = .5 m Case 6 6 6 6 6 6 6 Table 3.6. Case 6 Speed Study CFD Results Efficiency 𝑈 𝑚2 Run 𝐻(m) 𝑏(m) 𝑒̃ ( 𝑠2 ) 𝜂 RPM (m/s) b 3.79 0.11 33.37 89.22 200 10.47 d 3.79 0.14 33.65 90.02 170 8.90 f 3.78 0.18 33.70 90.21 140 7.33 h 3.78 0.22 33.44 89.70 110 5.76 j 3.79 0.26 32.61 87.25 80 4.19 l 3.75 0.08 32.12 86.74 250 13.09 CORD. 3.69 0.06 25.08 68.84 419 21.94 𝑊̇ (kW) 𝜙 0.79 0.93 1.13 1.43 1.97 0.63 0.38 Ω𝑠 2.82 2.39 1.98 1.55 1.13 3.55 6.02 𝜓 0.30 0.42 0.63 1.01 1.86 0.19 0.05 Peak efficiency was observed at a specific speed of 2.0, RPM of 140. Decreased efficiency was measured at rotational speeds above 170 RPM. 3.2. Discussion of Cordier Recommendations and Location on Diagram The specific speed (and rotational speed) recommended by the Cordier line are greater than the speed at which highest efficiency was measured in the CFD simulations. The efficiency at the Cordier-recommended speed was lower for each case tested in the Speed Study; however cases 3 and 6, which had the largest flow rate, had the smallest decrease in efficiency as speed increased. The efficiency obtained in CFD was graphed against the specific speed for each case in the Speed Study, shown in Figure 3.1 below. 86 136.83 137.97 138.17 137.12 133.70 131.68 131.68 Speed Study: Specific Speed Ωs vs. Efficiency 95.00 90.00 85.00 Efficiency 80.00 Case 1 75.00 Case 2 70.00 Case 3 Case 4 65.00 Case 5 60.00 Case 6 55.00 50.00 0.00 1.00 2.00 3.00 4.00 5.00 6.00 Specific Speed Ωs Figure 3.1 Specific speed 𝛺𝑠 Vs. Efficiency of Speed Study designs Efficiency at the specific speeds recommended by the Cordier line was lower than the efficiency at lower rotational speeds. Highest values of efficiencies were measured at specific speeds Ω𝑠 between 1.4 and 1.9, lower than the values of specific speed which were recommended by the Cordier line, between 5.5 and 6.1. The highest efficiency was measured for the designs which ran at speeds between 12 and 33% of the Cordier-recommended speed. The RPM, specific speed, and specific diameter where highest efficiency was measured in CFD for each case was logged, and compared to efficiency at the RPM and specific speed recommended by the Cordier line, recorded below in Table 3.7. 87 Table 3.7. Performance of Cordier Recommendations of Cases in Speed Study Case Units 1 2 𝑯 𝑚 1.85 1.88 𝒓𝒔 𝑚 0.2 𝒓𝒉 𝑚 𝑚3 𝑠 𝑽̇ 𝑫𝒔 RPM at highest 𝜼 𝛀𝒔 (highest 𝜼) Highest 𝜼 from Speed Study 3 4 5 6 1.91 3.70 3.70 3.70 0.3 0.5 0.2 0.3 0.5 .12 .18 .3 .12 .18 .3 𝑅𝑃𝑀 𝑟𝑎𝑑 0.46 1.218 170 1.37 1.05 1.214 140 1.69 2.85 1.232 90 1.77 0.60 1.268 300 1.65 1.50 1.202 90 0.78 4.10 1.212 140 2.01 % 80.58 85.49 89.76 81.99 85.96 90.21 6.1 704 22 52.44 6.0 419 22 68.84 Cordier Recommendations: Changing Speed 𝛀𝒔 (Cordier) RPM (Cordier) U (Cordier) 𝜼 at Cordier point 𝑟𝑎𝑑 𝑅𝑃𝑀 𝑚/𝑠 % 5.9 737 15 34.87 6.0 497 16 52.57 5.8 295 15 68.15 5.5 997 21 37.98 The below Figure 3.2 shows the six highest efficiency cases from the first study placed on the Cordier diagram from Wright [18]. Figure 3.2. Location of peak efficiency points for each case on 𝛺𝑠 vs 𝐷𝑠 diagram with Cordier line 88 Although the designs have specific diameter which the literature would recommend using an axial turbine, the designs with highest efficiency measured in the CFD study were located below the Cordier line, at rotational speeds less than half of that recommended by the Cordier line. The peak efficiency designs operated at specific speeds which were close to the specific speeds at which radial turbines typically operate. The specific diameter of the designs suggest the designs fall into Wright’s region A on the Cordier diagram, typically populated by axial turbines. The specific speeds of the peak efficiency designs, however, fit with Wright’s region F on the Cordier diagram, where high pressure blowers, centrifugal compressors, and high head liquid pumps are typically used. 3.3. Discussion of Speed Study Results 3.3.1. Overall Summary For each of the designs simulated in the Speed Study, specific diameter was relatively constant, and had a value of 𝐷𝑠 between 1.20 and 1.27. The Speed Study analysis used three sets of hub and shroud radii and blade thickness, with two values of total available head. Cases 1 and 4, 2 and 5, and 3 and 6 had the similar geometry, and had similar efficiencies, although the higher head designs (4 meters) were slightly higher efficiency than the designs for 2 meters. For each of the cases it was observed that as specific speed Ω𝑠 was increased past 2, turbine efficiency decreased. This is shown in Figure 3.3. As rotational speed was changed, tip speed 𝑈 changed, and flow coefficient 𝜑 changed. The below Figure 3.4, which shows efficiency graphed vs. flow coefficient, designs which had flow coefficients between 1.1 and 1.6 had the highest efficiency measured in simulations. Efficiency was lowest in the designs with the lowest flow coefficients, where rotational speed was high. 89 95.00 Speed Study: Flow Coefficient 𝜑 vs. Efficiency 90.00 Efficiency 85.00 80.00 Case 1 75.00 Case 2 70.00 Case 3 Case 4 65.00 Case 5 60.00 Case 6 55.00 50.00 0.00 0.50 1.00 1.50 2.00 2.50 Flow Coefficient 𝜑 Figure 3.3 Flow coefficient vs. Efficiency of Speed Study designs Designs which produced the highest efficiency had blade loading coefficients between 0.63 and 1.17. This is shown below in Figure 3.4. Efficiency was lowest for the designs with the lowest head coefficients, the designs which ran with Ωs above 2.5. Speed study: Blade Loading Coefficient 𝜓 vs. Efficiency 95.00 90.00 Efficiency 85.00 80.00 Case 1 75.00 Case 2 70.00 Case 3 65.00 Case 4 60.00 Case 5 Case 6 55.00 50.00 0.00 0.50 1.00 Blade Loading Coefficient 𝜓 1.50 2.00 Figure 3.4. Blade loading coefficient vs. Efficiency of Speed Study designs 90 3.3.2. Flow Visualization and Vectors The simulations of designs with high rotational speed predicted flows with higher relative flow speed. The higher speed designs also experienced larger flow velocity gradients at the leading and trailing edges. Towards the trailing edge on the pressure side, the flow accelerated and pressure decreased. This effect was more severe for higher rotational speed designs. Relative velocity vector diagrams were generated for each design, and show the velocity gradients at the leading and trailing edges. Figure 3.5 shows the flow of three designs in case 1, Figure 3.6 shows the flow of three designs in case 2, Figure 3.7 shows the flow of three designs in case 3, Figure 3.8 shows the flow of three designs in case 4, Figure 3.9 shows the flow of three designs in case 5, Figure 3.10 shows the flow of three designs in case 6. 1j Ω𝑠 = 1.37 𝜂 =78.4% 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 1b Ω𝑠 = 2.8 𝜂 = 77.1% Figure 3.5. Relative velocity vectors for case 1: Span 50% 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% 2d Ω𝑠 = 3.4 𝜂 = 78.7% Figure 3.6. Relative velocity vectors for case 2: Span 50% 91 2l Ω𝑠 = 1.7 𝜂 = 85.5% 3CORD Ω𝑠 = 5.9 𝜂 = 68.2% 3n Ω𝑠 = 1.8 𝜂 = 89.8% 3f Ω𝑠 = 4.9 𝜂 = 82.6% Figure 3.7. Relative velocity vectors for case 3: Span 50% 4CORD Ω𝑠 = 5.4 𝜂 = 37.9% 4h Ω𝑠 = 1.6 𝜂 = 81.9% 4b Ω𝑠 = 2.4 𝜂 = 79.6% Figure 3.8. Relative velocity vectors for case 4: Span 50% 5CORD Ω𝑠 = 6.1 𝜂 = 52.4% 5b Ω𝑠 = 3.9 𝜂 = 77.5% 5n Ω𝑠 = 1.6 𝜂 = 85.9% Figure 3.9. Relative velocity vectors for case 5: Span 50% 92 6CORD Ω𝑠 = 6.0 𝜂 = 68.8% 6b Ω𝑠 = 2.8 𝜂 = 89.2% 6f Ω𝑠 = 1.9 𝜂 = 90.2% Figure 3.10. Relative velocity vectors for case 6: Span 50% To show additional flow details, more detailed relative flow velocity vector diagrams were generated at the leading and trailing edges using the simulation results of each design. Figures 3.12, 3.13, and 3.14 show relative velocity vectors at leading edge for three designs within cases 2,1, and 3 respectively. At the leading edge, a region of flow deceleration was measured on the pressure side, located close to a region of rapid flow acceleration towards the suction side. The higher speed designs exhibited flows with a higher magnitude flow deceleration on the leading edge pressure side, and higher flow acceleration near the leading edge suction side. This same effect was observed for each of the cases simulated. 2d Ω𝑠 = 3.4 𝜂 = 78.7% 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% 2l Ω𝑠 = 1.7 𝜂 = 85.5% Figure 3.11. Relative velocity vectors at leading edge for case 2: 50% span 93 1b Ω𝑠 = 2.8 𝜂 = 77.1% 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 1j Ω𝑠 = 1.37 𝜂 = 80.6% Figure 3.12. Relative velocity vectors at leading edge for case 1: Span 50% 3CORD Ω𝑠 = 5.9 𝜂 = 68.2% 3f Ω𝑠 = 4.9 𝜂 = 82.6% 3n Ω𝑠 = 1.8 𝜂 = 89.8% Figure 3.13. Relative velocity vectors at leading edge for case 3: Span 50% Similarly, vector diagrams were produced for the trailing edge. The flow for the higher speed designs experienced more severe velocity gradients at and after the trailing edge. This relationship was observed for every case simulated in the Speed Study. This is shown below in Figure 3.14 for case 2, Figure 3.15 for case 1, and Figure 3.16 for case 3. 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% 2d Ω𝑠 = 3.4 𝜂 = 78.7% 2l Ω𝑠 = 1.7 𝜂 = 85.5% Figure 3.14. Relative velocity vectors at trailing edge for case 2: 50% span 94 1b Ω𝑠 = 2.8 𝜂 = 77.1% 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 1j Ω𝑠 = 1.37 𝜂 = 80.6% Figure 3.15. Relative velocity vectors at trailing edge for case 1: 50% span 3CORD Ω𝑠 = 5.9 𝜂 = 68.2% 3f Ω𝑠 = 4.9 𝜂 = 82.6% 3n Ω𝑠 = 1.8 𝜂 = 89.8% Figure 3.16. Relative velocity vectors at trailing edge for case 3: Span 50% 3.3.3 Trailing Edge Effects The above figures illustrate how varying rotational speed effected velocity gradients at the trailing edge. If the magnitude of the velocity gradients at the trailing edge are larger, additional viscous shear could occur, resulting in flow turbulence and entropy generation [35]. The designs with higher speed exhibited flows with more extreme velocity gradients at the trailing edge, which results in added viscous shear compared to the lower speed designs. The designs recommended by the Cordier line experienced the largest velocity gradients at the trailing edge, as well as the largest flow deceleration measured after the trailing edge. This effect was observed for every case simulated in the Speed Study. Flow will be more turbulent in areas where velocity gradients are more severe, and the higher speed designs in the Speed Study experienced greater velocity gradients at the trailing 95 edge. Turbulence kinetic energy was recorded to track the effect of increasing rotational speed on turbulence. Turbulence kinetic energy at and after the trailing edge was higher for the designs which operated at higher rotational speeds. Figure 3.17 shows a contour of turbulence kinetic energy for two designs 3Cord, 3n, and Figure 3.18 shows this for two designs in case 1, 1Cord and 1j. 3CORD Ω𝑠 = 5.9 𝜂 = 68.2% 3n Ω𝑠 = 1.8 𝜂 = 89.8% Figure 3.17. Turbulence kinetic energy contours for case 3: Span 50% 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 1j Ω𝑠 = 1.37 𝜂 = 80.6% 1l Ω𝑠 = 1.1 𝜂 = 77.5% Figure 3.18. Turbulence kinetic energy contours for case 1: Span 50% Figure 3.19 below shows the relative flow velocity for simulations of designs 1b ,1j and 1Cord of case 1 logged from 80% along the blade chord to the trailing edge. These figures illustrate the flow acceleration at the pressure side near the trailing edge, and rapid flow deceleration near the trailing edge on the suction side. As rotational speed was increased, flow 96 velocity at the pressure side trailing edge increased and velocity at the suction side trailing edge decreased. This was observed for every case simulated in the Speed Study. 1j Ω𝑠 = 1.37 𝜂 = 80.6% 𝑾 (𝒎/𝒔) 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 1b Ω𝑠 = 2.8 𝜂 = 77.1% Normalized position along blade Figure 3.19. Relative velocity of case 1 Speed Study designs near the trailing edge: Span 50% 3.3.4. Blade Loading Simulations showed that for the higher speed designs, the adverse pressure gradient on the suction side of the blades after the point of minimum pressure became more severe. Another adverse pressure gradient at the trailing edge at the suction side became more severe as speed increased. Adverse pressure gradients are correlated with boundary layer growth, entropy generation, and increased losses [35]. For each of the designs simulated in CFD, highest pressure was measured at the leading edge at the pressure side where flow was decelerated, the higher rotational speed designs demonstrated a higher pressure at this point. This effect was most extreme for the designs with rotational speeds recommended by the Cordier line. This point was located close to an area of lower pressure at the leading edge on the suction side, where flow accelerates. Simulations of the 97 designs with higher rotational speed predicted decreased pressure at this point compared to lower speed designs. This effect was observed for each case simulated in the Speed Study. For many designs, at this point pressure reached below vapor pressure. Towards the trailing edge on the pressure side, the flow accelerates and pressure decreases. This effect was magnified for higher rotational speed designs. The designs with Cordier speed experienced the most severe drop in pressure near the trailing edge on the pressure side. The area of low pressure at the trailing edge pressure side takes up a larger portion of the blade chord for the higher speed designs. Near the trailing edge on the suction side, pressure increases as the flow decelerates. Each of the higher rotational speed designs simulated recorded greater flow deceleration near the trailing edge on the suction side of the blades as the flow experienced an adverse pressure gradient. This is illustrated in blade loading diagrams for two of the cases, which chart pressure on the blades over the normalized position along the blade chord. Figure 3.20 shows design 2d, 2l, and the Cordier-recommended design for case 2, 2Cord, and Figure 3.21 shows design 1b, 1j, and the Cordier-recommended design for case 1, 1Cord. The adverse pressure gradients are highlighted with thicker lines. 98 2d Ω𝑠 = 3.4 𝜂 = 78.7% Pressure (Pa) 2l Ω𝑠 = 1.7 𝜂 = 85.5% 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% Normalized position along blade Figure 3.20. Blade loading diagrams for Speed Study case 2: Span 50% 1b Ω𝑠 = 2.8 𝜂 = 77.1% Pressure (Pa) 1j Ω𝑠 = 1.37 𝜂 = 80.6% 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% Normalized position along blade Figure 3.21. Blade loading diagrams for Speed Study case 1: Span 50% 99 Static pressure contours were produced for each design simulated, which display areas of increased pressure at the leading edge pressure side, decreased pressure at the leading edge on the suction side, and decreased pressure at the trailing edge pressure side. These figures illustrate that for the higher speed designs, the pressure gradients are of larger magnitude at each of these areas. This effect was observed for each case simulated in the Speed Study. Static pressure contours are shown for case 3 (Figure 3.22), case 1 (Figure 3.23), and case 3 (Figure 3.24) 2l Ω𝑠 = 1.7 𝜂 = 85.5% 2d Ω𝑠 = 3.4 𝜂 = 78.7% 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% Figure 3.22. Static pressure contours for Speed Study case 2: Span 50% 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 1b Ω𝑠 = 2.8 𝜂 = 77.1% 1j Ω𝑠 = 1.37 𝜂 = 80.6% Figure 3.23. Static pressure contours for Speed Study case 1: Span 50% 100 3CORD Ω𝑠 = 5.9 𝜂 = 68.2% 3n Ω𝑠 = 1.8 𝜂 = 89.8% 3f Ω𝑠 = 4.9 𝜂 = 82.6% Figure 3.24. Static pressure contours for Speed Study case 3: Span 50% Figure 3.25 and 3.27 below shows velocity triangles for two designs within the Speed Study, designs 1cord and 1j respectively, superimposed onto relative velocity vector contours. These figures illustrate how the relative velocities were higher for the designs with the Cordier recommended speeds. 101 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 𝑈 = 15.4 Figure 3.25. Velocity triangles with relative velocity contours for design 1CORD of the Speed Study: (Span 50%) 102 1j Ω𝑠 = 1.37 𝜂 = 80.6% 𝑈 = 3.6 Figure 3.26. Velocity triangles with relative velocity contours for design 1j of the Speed Study: (Span 50%) 3.3.5. Hub and Shroud Effects Decreased flow velocity was recorded near the hub and shroud surfaces for each design simulated. The higher speed designs saw relative velocity decrease the most near the hub and shroud. This effect was observed across each case in the Speed Study. Figure 3.27 shows the relative flow velocity 𝑊 over the blade span at the trailing edge for four designs in case 2 of the Speed Study, and Figure 3.28 shows this for four designs within case 1. 103 Span Normalized 2d Ω𝑠 = 3.4 𝜂 = 78.7% 𝑈 = 9.4 2n Ω𝑠 = 1.4 𝜂 = 84.6% 𝑈 = 3.8 2l Ω𝑠 = 1.7 𝜂 = 85.5% 𝑈 = 4.4 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% 𝑈 = 15.6 Span Normalized 𝑾 (𝒎/𝒔) Figure 3.27. Relative velocity of Case 2 Speed Study designs at the trailing edge along span 1j Ω𝑠 = 1.37 𝜂 = 80.6% 𝑈 = 3.6 1b Ω𝑠 = 2.8 𝜂 = 77.1% 𝑈 = 7.3 1l Ω𝑠 = 1.1 𝜂 = 77.5% 𝑈 = 2.9 1CORD Ω𝑠 = 5.5 𝜂 = 34.9% 𝑈 = 15.4 𝑾 (𝒎/𝒔) Figure 3.28. Relative velocity of Case 1 Speed Study designs at the trailing edge along span Figure 3.29 and Figure 3.30 display relative flow velocity for the designs at which highest efficiency was recorded in the Speed Study along the span of the blades, near the leading and trailing edges respectively. Decreased flow velocity close to the hub and shroud was recorded for each design simulated. 104 The designs with higher head (cases 4,5,6) exhibited flows with increased relative velocity compared to those with lower head (cases 1,2,3). The designs with the smallest diameter experienced the most flow deceleration near the hub and shroud at the trailing edge, these are designs in cases 1 and 4. It can be observed that lower velocity was recorded over a larger portion of the span at the trailing edge for the simulations of the designs with lower diameter, thus the boundary layer size was larger for the smaller diameter designs. The simulations of the designs with lower diameter predicted lower efficiency; this could be attributed in part to additional friction losses at the hub and shroud. Losses in the boundary layer are higher for turbine designs with increased surface area of the blades, as well as the hub and shroud surfaces Span Normalized relative to the volume of the fluid region. 2l Ω𝑠 = 1.7 𝜂 = 85.5% 𝑈 = 4.4 5n Ω𝑠 = 1.6 𝜂 = 85.9% 𝑈 = 5.9 3n Ω𝑠 = 1.8 𝜂 = 89.8% 𝑈 = 4.7 6f Ω𝑠 = 1.9 𝜂 = 90.2% 𝑈 = 7.3 1j Ω𝑠 = 1.37 𝜂 = 80.6% 𝑈 = 3.6 4h Ω𝑠 = 1.6 𝜂 = 81.9% 𝑈 = 6.3 𝑾 (𝒎/𝒔) Figure 3.29. Relative velocity of highest efficiency Speed Study designs at the leading edge along span 105 Span Normalized 4h Ω𝑠 = 1.6 𝜂 = 81.9% 𝑈 = 6.3 1j Ω𝑠 = 1.37 𝜂 = 80.6% 𝑈 = 3.6 2l Ω𝑠 = 1.7 𝜂 = 85.5% 𝑈 = 4.4 5n Ω𝑠 = 1.6 𝜂 = 85.9% 𝑈 = 5.9 3n Ω𝑠 = 1.8 𝜂 = 89.8% 𝑈 = 4.7 6f Ω𝑠 = 1.9 𝜂 = 90.2% 𝑈 = 7.3 𝑾 (𝒎/𝒔) Figure 3.30. Relative velocity of highest efficiency Speed Study designs at the trailing edge along span Below Figure 3.31 shows the turbulence kinetic energy 𝑘 at the trailing edge for the highest efficiency designs in the Speed Study. Turbulence kinetic energy of the flow was higher near the hub and shroud for the designs with higher head, and the designs with lower diameter (cases 1 and 4) showed the more turbulent region of the flow takes up a larger portion of the blade span. This could be attributed to additional friction losses at the hub and shroud in the designs with smaller diameter. 106 Span Normalized 3n Ω𝑠 = 1.8 𝜂 = 89.8% 1j Ω𝑠 = 1.37 𝜂 = 80.6% 4h Ω𝑠 = 1.6 𝜂 = 81.9% 2l Ω𝑠 = 1.7 𝜂 = 85.5% 6f Ω𝑠 = 1.9 𝜂 = 90.2% 5n Ω𝑠 = 1.6 𝜂 = 85.9% Turbulence kinetic energy 𝒌 Figure 3.31. Turbulence kinetic energy of highest efficiency Speed Study designs at the trailing edge along span 3.3.6. Work Extraction The simulations of each design predicted reduced work extraction near the hub and shroud. This is illustrated in the figures below, which display absolute circumferential flow velocity over the span of the blades at the trailing edge. Figure 3.32 shows absolute circumferential flow velocity for three designs in case 2 of the Speed Study, Figure 3.33 shows this for case 3. Designs which used the Cordier-recommended speed exhibited simulated flows with the largest decrease of circumferential flow velocity near the hub and shroud. This effect was observed for each case simulated in the Speed Study. 107 Span Normalized 2l Ω𝑠 = 1.7 𝜂 = 85.5% 𝑈 = 4.4 2CORD Ω𝑠 = 5.9 𝜂 = 52.6% 𝑈 = 15.6 2n Ω𝑠 = 1.4 𝜂 = 84.6% 𝑈 = 3.8 2d Ω𝑠 = 3.4 𝜂 = 78.7% 𝑈 = 9.4 𝑪𝒖 (𝒎/𝒔) Span Normalized Figure 3.32. Absolute circumferential flow velocity for designs in case 2 of Speed Study at the trailing edge along span 3CORD Ω𝑠 = 5.9 𝜂 = 68.2% 𝑈 = 15.4 3p Ω𝑠 = 1.2 𝜂 = 87.4% 𝑈 = 3.1 3f Ω𝑠 = 4.9 𝜂 = 82.6% 𝑈 = 13.1 3n Ω𝑠 = 1.8 𝜂 = 89.8% 𝑈 = 4.7 𝑪𝒖 (𝒎/𝒔) Figure 3.33. Absolute Circumferential flow velocity for designs in Case 3 of Speed Study at the trailing edge along span 108 For each design, 𝐶𝑢 reached a maximum near the trailing edge, and between the trailing edge and the exit the flow decelerates, and 𝐶𝑢 decreases. Figure 3.34 below shows average 𝐶𝑢 from annulus inlet to outlet for three designs within case 1 of the Speed Study. 𝑪𝒖 (𝒎/𝒔) 1l Ω𝑠 = 1.1 𝜂 = 77.5% 1j Ω𝑠 = 1.37 𝜂 = 80.6% 1b Ω𝑠 = 2.8 𝜂 = 77.1% Streamwise Location Figure 3.34. Average absolute circumferential flow velocity from inlet to outlet for designs case 1 of Speed Study Figure 3.35 below shows the absolute circumferential flow velocity for the designs with the highest predicted efficiency in the Speed Study along the blade span at the trailing edge. 𝐶𝑢 was higher near the hub than the shroud, this is because the relative flow angles were designed to have an equal value of 𝑈𝜃 𝐶𝑢𝑖 across the span. 109 Span Normalized 5n Ω𝑠 = 1.6 𝜂 = 85.9% 𝑈 = 5.9 4h Ω𝑠 = 1.6 𝜂 = 81.9% 𝑈 = 6.3 1j Ω𝑠 = 1.37 𝜂 = 80.6% 𝑈 = 3.6 2l Ω𝑠 = 1.7 𝜂 = 85.5% 𝑈 = 4.4 6f Ω𝑠 = 1.9 𝜂 = 90.2% 𝑈 = 7.3 3n Ω𝑠 = 1.8 𝜂 = 89.8% 𝑈 = 4.7 𝑪𝒖 (𝒎/𝒔) Figure 3.35. Absolute circumferential flow velocity for highest efficiency designs in Speed Study at the trailing edge along span 3.3.7. Cavitation It is important to track how low static pressure reaches when evaluating the performance of a hydraulic turbine design in CFD. If there are any areas where pressure reaches below the vapor pressure of water in simulation, cavitation could occur in a real-life application. Contour diagrams were produced for each design, recording areas where static pressure was close to or below the vapor pressure of water. Figure 3.36 shows the low static pressure contour for case 1, Figure 3.37 for case 2, Figure 3.38 for case 3, Figure 3.39 for case 4, Figure 3.40 for case 5, and Figure 3.41 for case 6. The figures show the simulations of designs at higher speeds exhibited flows with lower pressure at the leading edge on the suction side of the blades, and have added risk of experiencing cavitation. 110 1b Ω𝑠 = 2.8 𝜂 = 77.1% 1j Ω𝑠 = 1.37 𝜂 = 80.6% Figure 3.36. Areas of low static pressure for case 1 of Speed Study: 50% Span 2l Ω𝑠 = 1.7 𝜂 = 85.5% 2d Ω𝑠 = 3.4 𝜂 = 78.7% Figure 3.37. Areas of low static pressure for case 2 of Speed Study: 50% Span 3f Ω𝑠 = 4.9 𝜂 = 82.6% 3n Ω𝑠 = 1.8 𝜂 = 89.8% Figure 3.38. Areas of low static pressure for case 3 of Speed Study: 50% Span 111 4h Ω𝑠 = 1.6 𝜂 = 81.9% 4b Ω𝑠 = 2.4 𝜂 = 79.6% Figure 3.39. Areas of low static pressure for case 4 of Speed Study: 50% Span 5n Ω𝑠 = 1.6 𝜂 = 85.9% 5b Ω𝑠 = 3.9 𝜂 = 77.5% Figure 3.40. Areas of low static pressure for case 5 of Speed Study: 50% Span 6b Ω𝑠 = 2.8 𝜂 = 89.2% 6f Ω𝑠 = 1.9 𝜂 = 90.2% Figure 3.41. Areas of low static pressure for case 6 of Speed Study: 50% Span 112 3.4. Second Study: Blade Length Study The Blade Length Study consisted of simulations of designs with varying axial blade lengths at the rotational speeds at which highest efficiency was measured in the Speed Study. Different axial blade lengths will result in different values of Zweifel coefficient. The below tables show how varying Zweifel coefficient effected isentropic efficiency and trailing edge flow deviation. Table 3.8. Case 1 Blade Length Study Results Case 1 - 1j 𝐻 = 1.86 m 𝑚3 𝑉̇ = 0.46 𝑠 𝑟ℎ = .12 m 𝑟𝑠 = .2 m 𝑈 Case Run (𝑚/𝑠) 1j 1 3.56 1j 2 3.56 1j 3 3.56 1j 4 3.56 1j 5 3.56 1j 6 3.56 1j 7 3.56 1j 8 3.56 Efficiency 𝑚2 𝑒̃ ( 𝑠2 ) 𝑏 (m) 𝜂 RPM 14.45 0.162 78.40 170 14.69 0.122 79.04 170 14.51 0.090 79.72 170 14.64 0.070 79.70 170 14.34 0.050 78.21 170 14.13 0.190 77.43 170 14.83 0.080 80.06 170 14.78 0.096 79.96 170 Trailing edge deviation 𝛿 ′ (deg) 7.42 8.94 11.86 15.35 19.68 5.94 13.03 10.85 𝑊̇ (kW) 𝑍 0.48 0.63 0.85 1.09 1.49 0.41 0.96 0.80 6.65 6.76 6.68 6.73 6.59 6.50 6.82 6.80 Table 3.9. Case 2 Blade Length Study Results Case 2 - 2l 𝐻 = 1.85 m 𝑉̇ = 𝑚3 1.050 𝑠 𝑟ℎ = .2 m 𝑟𝑠 = .3 m 𝑈 Case Run (𝑚/𝑠) 2l 1 4.40 2l 2 4.40 2l 3 4.40 2l 4 4.40 2l 5 4.40 2l 6 4.40 2l 7 4.40 2l 8 4.40 𝑒̃ ( 𝑚2 ) 𝑠2 15.35 15.59 15.80 15.83 15.70 16.02 15.96 15.79 Efficiency 𝑏 (m) 𝜂 RPM 0.24 82.86 140 0.19 83.79 140 0.15 84.65 140 0.11 85.25 140 0.08 84.88 140 0.10 85.47 140 0.13 85.14 140 0.12 85.52 140 113 Trailing edge deviation 𝛿 ′ (deg) 5.71 7.15 8.99 12.61 17.10 14.57 10.34 10.63 𝑊̇ (kW) 𝑍 0.42 0.53 0.67 0.91 1.24 1.05 0.78 0.84 16.12 16.37 16.59 16.62 16.48 16.82 16.76 16.58 Table 3.10. Case 3 Blade Length Study Results Case 3 - 3n 𝐻 = 1.9 m 𝑚3 𝑉̇ = 2.85 𝑠 𝑟ℎ = .3 m 𝑟𝑠 = .5 m Case Run 3n 1 3n 2 3n 3 3n 4 3n 5 3n 6 3n 7 3n 8 3n 9 𝑈 (𝑚/𝑠) 4.71 4.71 4.71 4.71 4.71 4.71 4.71 4.71 4.71 𝑒̃ ( 𝑚2 ) 𝑠2 16.33 16.33 16.64 16.61 16.58 16.84 16.78 18.01 16.75 𝑏 (m) 0.40 0.35 0.30 0.25 0.20 0.15 0.13 0.11 0.18 Efficiency RPM 𝜂 86.79 90 86.76 90 88.11 90 88.57 90 89.30 90 89.77 90 89.71 90 87.58 90 89.94 90 Trailing edge deviation 𝛿 ′ (deg) 5.27 6.41 6.74 8.10 10.05 13.79 15.99 19.99 11.41 𝑊̇ (kW) 𝑍 0.40 0.46 0.53 0.64 0.79 1.05 1.20 1.42 0.90 46.54 46.55 47.44 47.35 47.26 47.99 47.81 51.33 47.73 Table 3.11. Case 4 Blade Length Study Results Case 4 - 4h 𝐻 = 3.69 m 𝑚3 𝑉̇ = 0.6 𝑠 𝑟ℎ = .12 m 𝑟𝑠 = .2 m Efficiency 𝑈 𝑚2 Case Run (𝑚/𝑠) 𝑒̃ ( 𝑠2 ) 𝑏 (m) 𝜂 RPM 4h 1 6.28 29.11 0.16 79.40 300 4h 2 6.28 29.33 0.14 80.17 300 4h 3 6.28 29.06 0.18 79.02 300 4h 4 6.28 29.73 0.11 81.15 300 4h 5 6.28 30.04 0.08 81.86 300 4h 6 6.28 29.60 0.06 81.99 300 4h 7 6.28 30.14 0.09 81.69 300 4h 8 6.28 30.14 0.05 81.77 300 4h 9 6.28 30.05 0.07 82.01 300 114 Trailing edge deviation 𝛿 ′ (deg) 5.45 6.25 4.99 7.76 10.77 14.63 9.52 15.88 12.54 𝑊̇ (kW) 𝑍 0.40 0.46 0.36 0.58 0.80 1.06 0.71 1.26 0.91 17.47 17.60 17.44 17.84 18.02 17.76 18.08 18.08 18.03 Table 3.12. Case 5 Blade Length Study Results Case 5 – 5n 𝐻 = 3.73 m 𝑚3 𝑉̇ = 1.5 𝑠 𝑟ℎ = .2 m 𝑟𝑠 = .3 m Case 6 6h 𝐻 = 3.78 m 𝑚3 𝑉̇ = 4.1 𝑠 𝑟ℎ = .3 m 𝑟𝑠 = .5 m Case Run 5n 1 5n 2 5n 3 5n 4 5n 5 5n 6 5n 7 5n 8 5n 9 𝑈 (𝑚/𝑠) 5.97 5.97 5.97 5.97 5.97 5.97 5.97 5.97 5.97 𝑒̃ ( 𝑚2 ) 𝑠2 30.85 30.48 31.07 31.05 31.41 31.48 31.40 31.55 31.49 𝑏 (m) 0.24 0.27 0.20 0.16 0.12 0.10 0.08 0.11 0.11 Efficiency RPM 𝜂 83.68 190 83.03 190 84.38 190 84.93 190 85.77 190 86.04 190 85.57 190 85.86 190 85.97 190 Trailing edge deviation 𝛿 ′ (deg) 6.08 5.45 7.16 8.80 11.45 13.89 17.90 12.61 10.66 Table 3.13. Case 6 Blade Length Study Results Trailing edge Efficiency deviation 𝑈 𝑏 𝑚2 Case Run (𝑚/𝑠) 𝑒̃ ( 𝑠2 ) (m) 𝜂 RPM 𝛿 ′ (deg) 6h 1 5.76 32.47 0.40 87.06 110 5.97 6h 2 5.76 32.39 0.44 86.72 110 5.52 6h 3 5.76 32.22 0.47 86.36 110 5.21 6h 4 5.76 31.98 0.50 85.97 110 4.93 6h 5 5.76 32.44 0.37 87.26 110 6.40 6h 6 5.76 32.68 0.33 87.92 110 7.07 6h 7 5.76 32.97 0.29 88.21 110 8.05 6h 8 5.76 33.10 0.25 88.76 110 9.22 6h 9 5.76 33.24 0.21 89.20 110 10.78 6h 10 5.76 33.52 0.16 89.58 110 14.19 6h 11 5.76 32.98 0.13 89.02 110 17.37 6h 12 5.76 33.36 0.20 89.81 110 10.97 𝑊̇ (kW) 𝑍 0.44 0.39 0.52 0.65 0.87 1.03 1.28 0.94 0.80 46.27 45.72 46.60 46.58 47.12 47.22 47.11 47.32 47.23 𝑊̇ (kW) 𝑍 0.44 0.41 0.38 0.36 0.48 0.54 0.61 0.70 0.83 1.08 1.31 0.90 133.14 132.78 132.10 131.13 132.99 134.00 135.17 135.73 136.30 137.45 135.23 136.78 3.5. Blade Length Study Results Discussion 3.5.1. Overall Summary For designs with long blade lengths Zweifel coefficient is small. These designs demonstrated lower efficiency in the CFD simulations than those with shorter blade lengths. Dixon suggested this is due to increased friction losses [4]. The below Figure 3.42 shows 115 efficiency obtained at differing values of Zweifel coefficient for the simulations conducted as a part of the Blade Length Study. At Zwefiel coefficients between 0.8 and 1.1 efficiency was highest in simulations, and stayed relatively constant across designs in that range. At Zwefiel coefficients below 0.8 predicted efficiency was lower as friction losses increased. Designs with Zweifel coefficients above 1.1 experienced decreased efficiency in simulations. Blade Length Study: 𝑍 vs Efficiency 91.00 89.00 Efficiency 87.00 Series1 85.00 Series2 83.00 Series3 81.00 Series4 79.00 Series5 Series6 77.00 75.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 Figure 3.42. Zweifel coefficient vs. Efficiency for Blade Length Study Flow does not follow the blade shape perfectly, and as blades are made shorter axially the blades must be made with sharper turning to obtain the same change in absolute circumferential flow speed and energy extraction. As the blade length was decreased and Zweifel coefficient increased, the trailing edge flow deviation 𝛿′ increased. In the Blade Length Study, a correlation between the Zweifel coefficient and trailing edge flow deviation was observed, this is shown below in Figure 3.43. 116 Trailing edge Deviation (deg) Blade Length Study: 𝑍 vs trailing edge deviation 𝛿′ 25.00 20.00 Series1 15.00 Series2 Series3 10.00 Series4 5.00 Series5 Series6 0.00 0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 Zweifel Coefficient 𝑍 Figure 3.43. Zweifel Coefficient vs. Flow deviation at trailing edge for Blade Length Study For the Blade Length Study simulations of turbines with both long axial blade lengths and short blade lengths were conducted, and it was observed that efficiency was highest for the simulations of designs with Zweifel coefficients between .8 and 1.1. With Zweifel coefficients below 0.8, efficiency decreased. This trend is illustrated in Figure 3.44, where the difference of the peak efficiency measured in the Speed Study and efficiency of the design simulated is shown vs. the Zweifel Coefficient for each case. 117 Peak efficiency-Efficiency simulated 4.50 Blade Length Study: Z vs Efficiency Decrease 4.00 3.50 3.00 Series1 2.50 Series2 2.00 Series3 1.50 Series4 1.00 Series5 0.50 Series6 0.00 0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 Zweifel Coefficient 𝑍 Figure 3.44. Difference of the peak efficiency and efficiency of the design vs. the Zweifel coefficient The decrease in predicted efficiency resulting from extending the blades longer than aerodynamically necessary can be considered modest, designs with Zweifel Coefficients near .6(±.05) predicted a decrease in efficiency no larger than 1.7% compared to the peak efficiency point. Designs with a Zweifel coefficient near .5(±.05) exhibited a maximum efficiency decrease of 2.5% compared to the peak efficiency point. Blades for water turbines experience high loads, and blades for water turbines often require large blade chords as longer blades can demonstrate improved structural performance, as distributing the force over a larger area decreases stresses on the blades [4]. Structural simulations were not done as a part of this study, however if FEA analysis suggested the designs with higher Zweifel coefficients would fail, longer blades could be employed to improve strength. 3.5.2. Flow Visualization and Vectors The relative velocity vector diagrams shown in the below figures at span 50%, show there are two areas where velocity gradients were had the highest magnitude, at the suction side leading edge, and on the pressure side at the trailing edge. These large changes in flow velocity 118 over short distances can lead to shear stresses and losses [35]. A wake was observed after the blade trailing edge where velocity is lower. The designs which had higher Zweifel coefficient had greater magnitude velocity gradients at the leading and trailing edges, and a larger wake size. Relative velocity vector contours were produced for each design in the Blade Length Study, Figure 3.45 shows relative velocity contours with vectors for case 1, Figure 3.46 for case 2, Figure 3.47 for case 3, Figure 3.48 for case 4, Figure 3.39 for case 5, and Figure 3.40 for case 6. 1j Z = 0.48 𝜂 = 78.4% 1j8 Z = 0.8 1j8 𝜂 = 79.9% 1j5 Z = 1.49 𝜂 = 78.2% Figure 3.45. Relative velocity vectors for case 1 of Blade Length Study: Span 50% 2l Z = 0.42 𝜂 = 82.9% 2l8 Z = 0.84 𝜂 = 85.5% 2l5 Z = 1.24 𝜂 = 84.9% Figure 3.46. Relative velocity vectors for case 2 of Blade Length Study: Span 50% 119 3n Z = 0.40 3n 𝜂 = 86.8% 3n6 Z=1.05 𝜂 = 89.8% Figure 3.47. Relative velocity vectors for case 3 of Blade Length Study: Span 50% 4h Z = 0.40 𝜂 = 79.4% 4h9 Z=0.91 𝜂 = 82.0% 4h6 Z=1.06 𝜂 = 81.9% Figure 3.48. Relative velocity vectors for case 4 of Blade Length Study: Span 50% 5n Z = 0.44 𝜂 = 83.7% 5n6 Z = 1.03 𝜂 = 86.0% Figure 3.39. Relative velocity vectors for case 5 of Blade Length Study: Span 50% 120 6h10 Z =1.08 𝜂 = 89.6% 6h12 Z =0.9 𝜂 = 89.9% 6h Z = .44 6h 𝜂 = 87.1% Figure 3.40. Relative velocity vectors for case 6 of Blade Length Study: Span 50% For any turbine the flow velocity is lower on the pressure side, and the velocity is higher on the suction side. For the designs with shorter blades, the simulated flow accelerated over a shorter distance. For the designs with larger Zweifel coefficients simulated in the Blade Length Study, the relative velocity was higher at the suction side, and the relative velocity at the pressure side was lower compared to designs with lower Zweifel coefficients. To show more detail into the flow features, vector diagrams were produced at the leading and trailing edges, vectors at the leading edge are shown below in Figure 3.41 for two designs in case 5. 5n6 Z = 1.03 𝜂 = 86.0% 5n Z = 0.44 5n 𝜂 = 83.7% Figure 3.41. Relative velocity vectors at leading edge for case 5 of Blade Length Study (Span 50%) As Zweifel coefficient was increased, flow acceleration at the leading edge suction side increased. This is shown in Figure 3.41 above, and is also seen below in Figure 3.42 for case 3, and Figure 3.43 for case 1. For each design simulated in the Blade Length Study, this effect was observed. 121 3n Z = 0.40 𝜂 = 86.8% 3n6 Z=1.05 𝜂 = 89.8% Figure 3.42. Relative velocity vectors at leading edge for case 3 of Blade Length Study (Span 50%) 1j5 Z = 1.49 𝜂 = 78.2% 1j Z = 0.48 𝜂 = 78.4% Figure 3.43. Relative velocity vectors at leading edge for case 1 of Blade Length Study (Span 50%) The below figures show the velocity vectors at the trailing edge for designs in cases 1, 5 and 3, Figures 3.55, 3.56, and 3.57 respectively. 5n Z = 0.44 𝜂 = 83.7% 5n4 Z=0.65 𝜂 = 84.9% 5n6 Z = 1.03 𝜂 = 86.0% Figure 3.44. Relative velocity vectors at trailing edge for case 5 of Blade Length Study (Span 50%) 122 3n Z = 0.40 𝜂 = 86.8% 3n6 Z=1.05 𝜂 = 89.8% Figure 3.45. Relative velocity vectors at trailing edge for case 3 of Blade Length Study (Span 50%) 1j8 Z = 0.8 𝜂 = 79.9% 1j5 Z = 1.49 𝜂 = 78.2% 1j Z = 0.48 𝜂 = 78.4% 1j5 Z = 1.49 𝜂 = 78.2% Figure 3.46. Relative velocity vectors at trailing edge for case 1 of Blade Length Study (Span 50%) 3.5.3. Trailing Edge Effects The above figures illustrate that as Zweifel coefficient was increased, flow acceleration near the trailing edge on the pressure side of the flow increased. On the suction side as flow travels towards the trailing edge, flow decelerates and experienced an adverse pressure gradient, and the boundary layer along this surface became larger. The designs with higher Zweifel coefficient experienced more flow deceleration towards the trailing edge on the suction side, velocity gradients at the leading and trailing edges, and larger size boundary layers. As Zweifel 123 coefficient was increased the wake left by the trailing edge became larger, and shifted closer toward the suction side. These effects can result in additional viscous shear, turbulence, entropy production, and losses at the trailing edge. This effect was observed for each of the designs that were simulated as a part of the Blade Length Study. Figure 3.47 below shows the magnitude of the relative flow velocity near the blade surface for designs 3n,3n3, and 3n6 of case 3, from 80% along the blade chord to the trailing edge. Each design experienced flow acceleration at the pressure side near the trailing edge. The figure shows the magnitude of this flow acceleration was highest for the designs with the shortest blades. The designs with larger Zweifel coefficients had greater velocity gradients over a smaller axial distance. This suggests the shear stresses and entropy production in this region are higher. This effect was observed for each design simulated as a part of the Blade Length Study. 𝑾 (𝒎/𝒔) 5n6 Z = 1.03 𝜂 = 86.0% 5n4 Z=0.65 𝜂 = 84.9% 5n Z = 0.44 𝜂 = 83.7% Normalized position along blade Figure 3.47. Relative velocity near trailing edge of case 3 and 5 of Blade Length Study Figure 3.48 below shows a contour of the turbulence kinetic energy two designs in case 1 of the Blade Length Study, Figure 3.49 shows two designs in case 3 and Figure 3.50 shows two 124 designs in case 5. These figures illustrate the increased turbulence at the trailing edge for the shorter blade length designs. 1j5 Z = 1.49 𝜂 = 78.2% 1j8 Z = 0.8 𝜂 = 79.9% Figure 3.48. Turbulence kinetic energy contour for cases 1j5 and 1j8: Span 50% 3n6 Z=1.05 3n6 𝜂 = 89.8% 3n Z = 0.40 𝜂 = 86.8% Figure 3.49. Turbulence kinetic energy contour for designs 3n and 3n6: Span 50% 5n Z = 0.44 𝜂 = 83.7% 5n6 Z = 1.03 𝜂 = 86.0% Figure 3.50. Turbulence kinetic energy contour for designs 5n and 5n6: Span 50% 125 Figure 3.51 below shows exit loss coefficient graphed vs Zweifel coefficient for each design in the Blade Length Study. For Zweifel coefficients under .8, the exit loss coefficient 𝜉2−𝑐 was relatively constant, however as Z was increased above 0.8, the exit loss coefficient increased, which is likely the result of increased losses resulting from the trailing edge wake. Blade length study: Z vs Exit Loss Coeff 𝝃𝟐−𝒄 0.070 Exit Loss Coefficient 0.060 0.050 Series1 0.040 Series2 Series3 0.030 Series4 0.020 Series5 Series6 0.010 0.000 0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 Zweifel Coefficient Z Figure 3.51. Zweifel Coefficeint vs Exit Loss Coefficient for Blade Length Study Turbines are usually designed to have a thin trailing edge. Blade profiles with high Zweifel coefficient require careful design and iteration to reduce trailing edge losses and reduce the risk of flow separation. Performance of the designs investigated in this work could be improved by tapering the trailing edge, especially those with high Zweifel coefficients, to reduce the trailing edge losses and wake size after the trailing edge. 3.5.4. Blade Loading With shorter blades, total pressure must decrease over a shorter axial distance. This can lead to more flow deviation and velocity gradients at the leading and trailing edges, which can lead to additional losses. The designs with higher values of 𝑍 exhibited flows with lower 126 pressure on the suction side of the blade, and higher pressure on the pressure side. With shorter axial length, the pressure difference from pressure to suction side needs to be larger to keep power constant, as the area on which the pressure acts is smaller, and the same force must be maintained to keep power constant. Simulations showed that as Zweifel coefficient was increased by decreasing blade length, an adverse pressure gradient developed on the suction side of the blades after the point of minimum pressure. Another adverse pressure gradient at the trailing edge at the suction side became more pronounced as Zweifel coefficient was increased. Adverse pressure gradients are correlated with boundary layer growth, entropy generation, and increased losses. Below Figure 3.52 shows the pressure vs. blade position (blade loading) charts for case 3 at three Zweifel coefficients, at 50% span. Figure 3.53 shows static pressure contours for the three designs within case 3 of the Blade Length Study. Figure 3.54 and 3.66 show the blade loading chart and static pressure contours respectively for three designs within case 5. The shorter blade length designs experienced flows with lower pressure on the suction side of the blades, the figures illustrate this effect. For each design simulated in this work, towards the trailing edge an area of low pressure and flow acceleration on the pressure side of the blade was observed, located close next to an area of flow deceleration, an adverse pressure gradient, at the suction side trailing edge. As the blade length was decreased this effect was magnified, and the region was observed over a larger portion of the distance along the blade. 127 3n Z = .40 3n3 Z = .53 η=88.1% 3n6 Z = 1.05 η=89.77% Pressure (Pa) η=86.8% Normalized position along blade Figure 3.52. Blade loading diagrams for case 3 of Blade Length Study: Span 50% 3n Z = .40 𝜂 = 86.8% 3n3 Z = .53 𝜂 = 88.1% 3n6 Z = 1.05 𝜂 = 89.77% Figure 3.53. Static Pressure Distributions for Case 3 of Blade Length Study (Span 50%) 128 5n Z = .44 5n4 Z = .65 η=83.7% 5n6 Z = .1.03 η=86.0% Pressure (Pa) η=84.9% Normalized position along blade Figure 3.54. Blade loading diagrams for case 5 of Blade Length Study: Span 50% 5n Z = .44 𝜂 = 83.7% 5n4 Z = ..65 5n4 𝜂 = 84.9% 5n6 Z = .1.03 η=86.0% Figure 3.55. Static Pressure Distributions for Case 5 of Blade Length Study (Span 50%) Figures 3.67 and 3.68 below show velocity triangles superimposed over velocity vector contours and blade loading diagrams produced in CFD-Post for two designs in case 2, designs 2l and 2l8. The figure illustrates how as Zweifel coefficient increases, velocity on the suction side of the blades increases, and velocity on the pressure side of the blades decreases. 129 Pressure (Pa) Figure 3.56.Velocity triangles superimposed over velocity vector contours for design 2l of the Blade Length Study at leading and trailing edge (span 50%) 130 Pressure (Pa) Figure 3.57.Velocity triangles superimposed over velocity vector contours for design 2l8 of the Blade Length Study at leading and trailing edges (span 50%) 3.5.5. Hub and Shroud Effects Literature suggests that for the designs with higher Zweifel coefficient, losses via friction from the blade surfaces, hub and shroud are reduced, however the losses resulting from diffusion and shear stresses at the trailing edge are higher [24] [29]. The results obtained in the Blade Length Study correlate with these findings. 131 To investigate the effects of friction resulting from the hub and shroud surfaces, Charts were generated which tracked the relative velocity of the flow along the span of the blades. Figure 3.58 below shows relative flow velocity graphed along the span of the blades at the leading edge of the turbine for case 3, and Figure 3.59 shows the relative flow velocity graphed along the span at the trailing edge. Decreased relative flow velocity was exhibited near the hub and shroud for each design tested, indicating friction effects. The designs with higher Zweifel coefficient experienced a larger decrease in relative velocity at the hub and shroud as the flow traveled from leading to trailing edge. Although the velocity distribution at the leading edge was similar for each of the designs, the designs with lower Zweifel coefficient experienced decreased relative flow velocity at the trailing edge, especially near the hub and shroud. This suggests the lower Z designs experienced additional friction resulting from the hub and shroud surfaces, compared to the higher Z designs. Span Normalized 6h12 Z =0.9 𝜂 = 89.9% 6h10 Z =1.08 𝜂 = 89.6% 6h Z = .44 𝜂 = 87.1% 6h4 Z = .36 𝜂 = 85.9% 𝑾(𝒎/𝒔) Figure 3.58. Relative velocity of case 6 for Blade Length Study at the leading edge along span 132 Span Normalized 6h12 Z =0.9 𝜂 = 89.9% 6h10 Z =1.08 𝜂 = 89.6% 6h Z = .44 𝜂 = 87.1% 6h4 Z = .36 𝜂 = 85.9% 𝑾(𝒎/𝒔) Figure 3.59. Relative velocity of case 6 for Blade Length Study at the trailing edge along span Profile loss coefficient 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 was graphed vs the Zweifel coefficient to track how Z effected losses from the leading to trailing edge. Figure 3.60 below shows that as Zweifel coefficient was increased from the lowest values to .8, 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 decreased. However for the Zweifel coefficients above .8, 𝜉𝑝𝑟𝑜𝑓𝑖𝑙𝑒 stayed relatively constant. 133 Blade length study: Z vs Profile Loss 𝝃𝒑𝒓𝒐𝒇𝒊𝒍𝒆 0.80 Profile loss Coefficent 0.75 0.70 Series1 Series2 0.65 Series3 Series4 0.60 Series5 Series6 0.55 0.50 0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 Zweifel Coefficent Z Figure 3.60. Zweifel coefficeint vs Profile loss coefficient for Blade Length Study 3.5.6. Work Extraction Average relative flow velocity was graphed from annulus inlet to annulus outlet. Figure 3.61 below shows average relative velocity from inlet to outlet for case 1, Figure 3.62 for case 6. The designs with higher efficiency had higher relative velocity at the trailing edge, however each design had equal relative velocities at the leading edge. As 𝑊𝑚 stays constant from inlet to outlet, the highest efficiency designs had the highest average relative circumferential velocity at the trailing edge, 𝑊𝑢2 . 134 𝑾(𝒎/𝒔) 1j8 Z = 0.8 𝜂 = 79.9% 1j Z = 0.48 𝜂 = 78.4% 1j5 Z = 1.49 𝜂 = 78.2% Streamwise Location Figure 3.61. Average relative velocity from inlet to outlet for case 1 for Blade Length Study 6h12 Z =0.9 𝜂 = 89.9% 𝑾(𝒎/𝒔) 6h10 Z =1.08 𝜂 = 89.6% 6h Z = .44 𝜂 = 87.1% 6h4 Z = .36 𝜂 = 85.9% Streamwise Location Figure 3.62. Average relative velocity from inlet to outlet for case 6 for Blade Length Study The simulations of each design produced results which predicted reduced work extraction near the hub and shroud for each design. This is illustrated in the figures below, which display 135 absolute circumferential flow velocity over the span at the trailing edge. Figure 3.63 below shows absolute circumferential velocity graphed along the span of the blades for three designs in case 6 of the Blade Length Study, at the trailing edge, and Figure 3.64 shows this for case 1. The designs with Zweifel coefficients between 0.8 and 1.1 exhibited the highest absolute circumferential velocity 𝐶𝑢 near the hub at the trailing edge. These designs had the highest change in angular momentum, and therefore work extraction, near the hub compared to those with Zweifel coefficients outside this range. Span Normalized 6h12 Z =0.9 𝜂 = 89.9% 6h10 Z =1.08 𝜂 = 89.6% 6h Z = .44 𝜂 = 87.1% 6h4 Z = .36 𝜂 = 85.9% 𝑪𝒖 (𝒎/𝒔) Figure 3.63. Absolute circumferential velocity for case 6 of the Blade Length Study at the trailing edge along span 136 Span Normalized 1j8 Z = 0.8 𝜂 = 79.9% 1j5 Z = 1.49 𝜂 = 78.2% 1j Z = 0.48 𝜂 = 78.4% 𝑪𝒖 (𝒎/𝒔) Figure 3.64. Absolute circumferential velocity for case 1 of the Blade Length Study at the trailing edge along span Graphs were generated which show 𝐶𝑢 from inlet to outlet for each design, Figure 3.65 below shows this for case 6, and Figure 3.66 for case 1. The designs which had the highest predicted efficiency had the highest circumferential velocity at the trailing edge (𝐶𝑢2 ). As absolute circumferential velocity at the leading edge, 𝐶𝑢1 = 0, this means the highest efficiency designs exhibited flows with the largest change in absolute circumferential velocity 𝐶𝑢2 − 𝐶𝑢1 . These designs had the most work extraction, they exhibited flows with the largest change in angular momentum. 137 6h10 Z =1.08 𝜂 = 89.6% 6h4 Z = .36 𝜂 = 85.9% 𝑪𝒖 (𝒎/𝒔) 6h12 Z =0.9 𝜂 = 89.9% 6h Z = .44 𝜂 = 87.1% Streamwise Location Figure 3.65. Average Absolute circumferential velocity from inlet to outlet of case 6 of Blade Length Study 𝑪𝒖 (𝒎/𝒔) 1j8 Z = 0.8 𝜂 = 79.9% 1j Z = 0.48 𝜂 = 78.4% 1j5 Z = 1.49 𝜂 = 78.2% Streamwise Location Figure 3.66. Average Absolute circumferential velocity from inlet to outlet of case 1 of Blade Length Study To help observe the effect of reducing Zweifel coefficient on the simulated work extraction, the product of absolute circumferential velocity 𝐶𝑢 and rotational velocity 𝑈𝜃 was 138 graphed over the span of the at the trailing edge for each design. Figure 3.67 below shows 𝑈𝜃 𝐶𝑢 graphed for three designs in case 1 of the Blade Length Study at the trailing edge of each turbine. The highest efficiency designs had the highest value of 𝑈𝜃 𝐶𝑢 at the trailing edge. As Zweifel coefficient was increased, work extraction 𝑈𝜃 𝐶𝑢 at the trailing edge increased. However as Zweifel coefficient was increased above 1.2, 𝑈𝜃 𝐶𝑢 at the trailing edge decreased. This effect can be seen below in Figure 3.67 below, although the same effect was observed for each case in the Blade Length Study. The designs were set up to have no pre-swirl, such that 𝐶𝑢1 = 0. Designs which had the highest 𝑈𝜃 𝐶𝑢 at the trailing edge had the highest efficiency, as these designs exhibited flows with the largest change in angular momentum, and thus the largest shaft Span Normalized power. 1j5 Z = 1.49 𝜂 = 78.2% 1j Z = 48 𝜂 = 78.4% 1j8 Z = 0.8 𝜂 = 79.9% 𝑼𝜃 𝑪𝒖 (𝒎𝟐 /𝒔𝟐 ) Figure 3.67. 𝑈𝜃 𝐶𝑢 for case 1 of the Blade Length Study at the trailing edge across span 3.5.7. Cavitation As blade length was decreased to lower Zweifel coefficient, areas of low static pressure developed on the suction side of the blades, which could lead to cavitation in a real-world application. The below figures show areas of low static pressure for different designs in the 139 Blade Length Study. Figure 3.68 shows the low static pressure contour for case 1, Figure 3.69 for case 2, Figure 3.70 for case 3, Figure 3.71 for case 4, Figure 3.72 for case 5, and Figure 3.73 for case 6. The results suggest that as Z is increased, the risk of cavitation became greater. 1j8 Z = 0.8 𝜂 = 79.9% 1j Z = 48 𝜂 = 78.4% 1j5 Z = 1.49 𝜂 = 78.2% Figure 3.68. Areas of low static pressure for case 1 of Blade Length Study (Span 50%) 2l Z = 0.42 𝜂 = 82.9% 2l8 Z = 0.84 𝜂 = 85.5% Figure 3.69. Areas of low static pressure for case 2 of Blade Length Study (Span 50%) 3n Z = .40 𝜂 = 86.8% 3n6 Z = 1.05 𝜂 = 89.77% Figure 3.70. Areas of low static pressure for case 3 of Blade Length Study (Span 50%) 140 4h8 Z=1.26 𝜂 = 81.8% 4h9 Z=0.91 𝜂 = 82.0% 4h Z = 0.40 𝜂 = 79.4% Figure 3.71. Areas of low static pressure for case 4 of Blade Length Study (Span 50%) 5n Z = .44 𝜂 = 83.7% 5n6 Z = 1.03 𝜂 = 86.0% Figure 3.72. Areas of low static pressure for case 5 of Blade Length Study (Span 50%) 6h12 Z =0.9 𝜂 = 89.9% 6h Z = .44 𝜂 = 87.1% 6h10 Z =1.08 𝜂 =Z89.6% 6h10 =1.08 𝜂 = 89.6% Figure 3.73. Areas of low static pressure for case 6 of Blade Length Study (Span 50%) 141 CHAPTER 4. CONCLUSIONS No flow separation was observed for any of the designs simulated, and after optimizing the rotational speed and blade length each set of specifications had a design with efficiency predicted via CFD over 79%. This is despite the fact the highest efficiency designs simulated were located below the Cordier Line. The designs with lower diameter and smaller crosssectional area showed significantly lower predicted efficiency, Cases 1 and 4 have a tip diameter of .2 meters, designs done for those cases had a peak efficiency of 80%, compared to 90% for cases 3 and 6, which had a tip diameter of .5 meters. Both hypotheses introduced in section 1.17 were confirmed in this work. First, the designs simulated with specific speed which place it on the Cordier line performed with lower efficiency than designs with specific speed lower than that recommended by the line. Second, peak efficiency in CFD was obtained in CFD for designs which had Zweifel coefficients between .8 and 1.1. 4.1. Speed Study and Cordier Line For the Speed Study, six sets of turbine specifications were used to produce turbine designs, which were tested in CFD. Designs were generated and simulated at differing rotational speeds. Highest efficiency was observed in the designs which ran at speeds at a range 12-33% that of the Cordier-recommended speed. Highest efficiencies were recorded when the designs were generated at specific speeds between 1.37 and 1.98, lower than the values of specific speed recommended by the Cordier line, between 5.5 and 6.1. Designs which were generated for higher rotational speeds experienced more severe adverse pressure gradients on the suction side of the blades in simulation, and more severe velocity gradients near the leading and trailing edges. Turbulence kinetic energy at the trailing 142 edge was higher for the designs which operated at higher rotational speeds. The simulations suggest that as rotational speed was increased, the risk of cavitation on the suction side increased. 4.2. Blade Length Study and Zweifel Coefficient For the Blade Length Study, the designs at which highest efficiency was measured in the Speed Study were simulated with different Zweifel coefficients. Zweifel coefficient was changed by adjusting axial blade length 𝑏. Lower efficiency was predicted for the designs with Zweifel coefficients less than 0.75. Simulated efficiency was highest in the range of 0.8 < 𝑍 < 1.1, within this range efficiency was relatively constant. With higher Zweifel coefficients, pressure on the suction side is reduced and pressure on the pressure side is increased. Designs which had higher Zweifel coefficients experienced more severe adverse pressure gradients on the suction side of the blades. The higher Zweifel coefficient designs experienced more losses at the trailing edge and had a larger wake after the trailing edge. As Zweifel coefficient was increased, the risk of cavitation on the suction side increased. Increased Zweifel coefficients were correlated with larger flow velocity gradients at the leading and trailing edges of the blades. Trailing edge flow deviation increased linearly as Zweifel coefficient was increased. Literature suggests that for the designs with higher Zweifel coefficient, losses via friction from the blade surfaces, hub and shroud are reduced, however the losses resulting from diffusion and shear stresses at the trailing edge are higher. The results obtained in the Blade Length Study are in line with this prediction. Profile loss coefficient was lowest for designs within 0.8 < 𝑍 < 1.1, however as Zweifel coefficient was increased above 0.8, loss coefficient of the exit region increased, the result of increased trailing edge losses. Turbine designs with Zweifel coefficients 143 below 0.8 were also tested, it was shown that longer blade lengths could be employed, although these designs exhibited more friction losses and lower predicted efficiency in simulations. 4.3. Next Steps of the Project Additional simulation work can be done to shed more light on the performance of CTPATs. The simulations done in this work demonstrate the efficacy of the designs; however the manufactured turbine will have real geometry different than the idealized 3-D modeled geometry simulated. The Woven Wheel turbomachinery manufacturing technique offers benefits, however additional work is required to mature the manufacturing process to the point where it could be used to produce turbine rotors within acceptable tolerances. At high Reynolds numbers the surface finish of the blades become more important to keep friction losses low [35]. Blade thickness was kept constant across each design within each set of specifications, referred to as a case. To achieve more complete dynamic similarity between each design in the Speed Study 𝑡 𝑏 could be kept constant across each simulation, but for some of the higher speed simulations axial blade length must be decreased to keep Zweifel coefficient at .8, many of the blade thicknesses would be under 1 millimeter, which would be more costly to manufacture and could experience structural failure. The designs tested in this analysis did not have inlet or outlet guide vanes. Further simulations would show how this affects the performance of the turbine at off-design points, at differing operating conditions. The simulations assumed a uniform laminar velocity profile at the turbine inlet. This is an idealized scenario, in a real world application flow would be more turbulent at the inlet and a nose cone/guide vanes would be required to guide the flow into the turbine similar to the one simulated in this study. Transient simulations would reveal additional details into the flow features. The speed and blade length studies could be repeated with different 144 specifications, altering volume flow rate, head, and turbine geometry. This could reveal additional insights into how the inputs effect the specific speed at which peak efficiency is measured. ANSYS CFX includes cavitation modeling as a part of the simulation package, designs of this type should be simulated with cavitation modeling to see where cavitation bubbles develop and collapse. The design could then be improved to reduce the risk of cavitation. Finite Element Analysis (FEA) simulations of the blades would validate the structural performance of the turbine designs. FEA can be used to find the points with the highest stresses, and the design can be changed to reduce stress in those areas. Performance would be improved by using blade profiles designed for water turbines rather than using constant thickness blade profiles with rounded leading and trailing edges. Trailing edge thickness could be reduced to reduce flow deviation, reduce trailing edge losses and flow deviation. 145 REFERENCES 146 REFERENCES [1] US. Department of Energy Wind and Water Power Technologies Office, "US. Department of Energy Wind and Water Power Technologies Office Funding in the United States: Hydropower Projects," 2012. [2] S. e. a. Sangal, "Review of Optimal Selection of Turbines for Hydroelectric Projects," International Journal of Emerging Technology and Advanced Engineering, vol. 3, no. 3, p. 426, March 2013. [3] P. Drtina and M. Sallaberger, "Hydraulic Turbines—Basic Principles and State-of-the-art Computational Fluid Dynamics Applications," in Proceedings of the Institute of Mechanical Engineers, 1999. [4] S. L. Dixon and C. A. Hall, Fluid Mechanics and Thermodynamics of Turbomachinery, Elsevier, 2010. [5] S. Adhau, "Economic Analysis and Application of Small Micro/Hydro Power Plants," in International Conference on Renewable Energies and Power Quality, April 2009. [6] B. Hadjerioua, Y. Wei and S.-C. Kao, "An Assesment of Energy Potential at Non-Powered Dams in the United States," April 2012. [7] B. Ogayar and P. G. Vidal, "Cost determination of the electro-mechanical equipment of a small hydro-power plant," Renewable Energy, vol. 34, pp. 6-13, 2009. [8] N. Mueller.United States Patent US 7938627 B2, 2011. [9] N. e. a. Mueller, "Low Cost Wound and Woven Turbomachinery Design," in Proceedings of the ASME Turbo Expo 2007, 2007. [10] Dupont, "Kevlar Aramid Fiber Technical Guide," 2017. [Online]. Available: http://www.dupont.com/content/dam/dupont/products-and-services/fabrics-fibers-andnonwovens/fibers/documents/Kevlar_Technical_Guide.pdf. [11] J. H. Horlock, Axial Flow Turbines, Huntington, New York: Robert E. Krieger Publishing Company, 1966. [12] S. A. Korpela, Principles of Turbomachinery, Hoboken, NJ: John Wiley and Sons, Inc., 2011. [13] O. Cordier, ""Similarity Considerations in Turbomachines"," VDI Reports, 1955. [14] T. Wright, Fluid Machinery: Performance, Analysis, and Design, CRC Press, 1999. 147 [15] O. E. Balje, Turbomachines: A Guide to Design, Selection and Theory, New York: John Wiley and Sons, 1981. [16] Andritz Hydro, "Compact Hydro," September 2014. [Online]. Available: https://www.andritz.com/resource/blob/33446/089d4393ed0b9ee30e2f9ffba115430d/hycompact-hydro-en-data.pdf. [17] P. Adhikari, U. Budhathoki, S. R. Timilsina, S. Manandhar and T. R. Bajracharya, "A Study on Developing Pico Propeller Turbine for Low Head Micro Hydropower Plants in Nepal," Journal of the Institute of Engineering, Nepal, vol. 9, no. 1, pp. 36-53, 2013. [18] S. F. Smith, "A Simple Correlation of Turbine Efficiency," Journal of the Royal Aeronautics Society, 1965. [19] A. F. El-Sayed, Aircraft Propulsion and Gas Turbine Engines, Boca Raton, FL: CRC Press, 2008. [20] J. D. Coull and H. P. Hodson, "Blade Loading and its Application In The Mean Line Design of Low Pressure Turbines," in Proceedings of the ASME Turbo Expo 2011, Vancouver, 2011. [21] S. C. Kacker and U. Okapuu, "A Mean Line Prediction Method for Axial Flow turbine Efficiency," Journal of Engineering Power, vol. 104, pp. 111-119, 1982. [22] M. T. Schobeiri, Turbomachinery Flow Physics and Dynamic Performance, New York: Springer, 2005. [23] O. Zweifel and B. &. C. Brown, "The Spacing of Turbomachine Blading, Especially with Large Angular Deflection," 1945. [24] E. Dick, Fundamentals of Turbomachines, Springer, 2015. [25] C. Prakash, D. G. Cherry, H. W. Shin, J. Machnaim, L. Dailey, R. Beacock, D. Halstead and A. R. Wadia, "Effect of Loading Level and Distribution on LPT Losses," in Proceedings of ASME Turbo Expo 2008, Berlin, Germany, 2008. [26] S. C. M. R. M. a. T. C. C. J. P. C. P. J. K. a. S. L. P. J. T. Schmitz, "Highly Loaded LowPressure Turbine: Design, Numberical, and Experimental Analysis," in Proceedings of the ASME 2010 Turbo Expo, Glasgow, Scotland, 2010. [27] H. Pfeil, Optimale Primarverluste in Axialgittern und Axialstufen von Stromungsmaschinen, VD-Forschungsheft, p. 535. [28] J. T. Schmitz, S. C. Morris, R. Ma, T. C. Corke, J. P. Clark, P. J. Koch and S. L. Puterbaugh, "Highly Loaded Low-Pressure Turbine: Design, Numerical, and Experimental Analysis," in Proceedings of the ASME Turbo Expo 2010, 2010. 148 [29] T. J. Praisner , E. A. Grover and D. C. Knezevici, "Toward the Expansion of Low-Pressure Turbine Airfoil Design Space," in Proceeedings of the ASME Turbo Expo 2008, Germany, 2008. [30] M. Ćosić, M. Dojčinovi and Z. Aćimović-Pavlovic, "Fabrication and Behaviour of AlSi/SiC Composite in Cavitation Conditions," International Journal of Cast Metals Research, vol. 27, 2014. [31] T. Yamatogi, H. Murayama, K. Uzawa, K. Kageyama and N. Watanabe, "Study on Cavitation Erosion of Composite Materials for Marine Propeller," in ICCM International Conferences on Composite Materials, 2009. [32] D. D. J, "Loss Mechanisms in Turbomachines," in International Gas Turbine and Aeroengine Congress and Exposition, Cincinatti, Ohio, 1993. [33] W. Fox, W. Pritchard and A. McDonald, Introduction to Fluid Mechanics, Seventh Edition ed., John Wiley and Sons, 2010. [34] Hydraulic Institute, Pipe Friction Manual, New York, 1954. [35] C. Canuto, M. Y. Hussani, A. Quarteroni and T. A. Zang, Spectral Methods: Evolution to Complex Geometries and Applications to Fluid Dynamics, Springer, 2007. [36] Voith Hydro, "Environmentally Friendly Turbine Design," May 2011. [Online]. Available: http://www.voith.com/corp-de/VH_Product_Brochure_Environmentally-friendly-turbinedesign_14_vvk_t3360e_en.pdf. [37] J. M. Becker, "Identifying the Effects on Fish of Changes in Water Pressure during Turbine Passage," Hydro Review, vol. 22, no. 5, pp. 1-4, September 2003. [38] G. Cada, "The Development of Advanced Hydroelectric Turbines to Improve Fish Passage Survival," Fisheries, p. 14, September 201. [39] S. Lieblein, F. C. Schwenk and R. L. Broderick, "Diffusion Factor for Estimating Losses and Limiting Blade Loadings in Axial Flow Compessor Blade Elements," National Advisory Committee for Aeronautics, Washington, 1953. [40] R. W. Johnson, The Handbook of Fluid Dynamics, Boca Raton, FL: CRC Press, 2000. [41] W. W. Peng, Fundamentals of Turbomachinery, Hoboken, NJ: John Wiley and Sons, 2007. [42] D. R. Durgaiah, Fluid Mechanics and Machinery, New Dehli: New Age International Ltd., 2002. 149