You are on page 1of 323

CFD Open Series - Patch 2.

20

Essentials of Turbo
Machinery in CFD
Edited by :
Ideen Sadrehaghighi, Ph.D.

Unsteady
Unsteady
Flow in
Flow in Axial
Radial
Turbomachines
Turbomachines
(ANSYS)
(ANSYS)
Blade
Interaction
(NUMECA)
Goldman
Annular
Turbine
Cascade
(NASA-TGRID)

ANNAPOLIS, MD
++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++
2

Table of Contents

1 Introduction ................................................................................................................................ 20
1.1 Preliminaries......................................................................................................................................................... 20
1.2 Full Engine Simulation Methodology .......................................................................................................... 21

2 Preliminary Concepts in Rotating Machinery ................................................................. 23


2.1 Vortex....................................................................................................................................................................... 23
2.1.1 Properties of Vortex Flow .................................................................................................................. 23
2.1.1.1 Vorticity ......................................................................................................................................... 23
2.1.1.2 Vortex Types ............................................................................................................................... 24
2.1.1.2.1 Rigid-Body Vortex ........................................................................................... 24
2.1.1.2.2 Irrotational Vortex .......................................................................................... 24
2.1.2 Vortex Geometry .................................................................................................................................... 24
2.1.3 Pressure in Vortex ................................................................................................................................. 25
2.2 Impeller ................................................................................................................................................................... 25
2.2.1 Types of Impeller ................................................................................................................................... 26
2.2.2 Flow Characteristics for Impeller.................................................................................................... 26
2.2.3 Mixing Tanks ............................................................................................................................................ 27
2.2.4 Axial Impellers ........................................................................................................................................ 28
2.2.5 Radial Impellers...................................................................................................................................... 28
2.2.6 Power Number for Impeller .............................................................................................................. 28
2.3 Pumps ...................................................................................................................................................................... 28
2.3.1 Types of Pumps....................................................................................................................................... 29
2.3.2 Axial-Flow Pumps vs. Centrifugal Pumps .................................................................................... 30
2.4 Some Physics on Rotating Disks Flow ........................................................................................................ 30
2.4.1 Experimental Set-Up............................................................................................................................. 30
2.4.1.1 Recirculating Flow .................................................................................................................... 31
2.4.1.2 Instability Flow Patterns ........................................................................................................ 31
2.5 Discussion on Effects Swept and Dihedral Blades ................................................................................. 33
2.6 Linear Cascade ..................................................................................................................................................... 34

3 Conservation of Angular Momentum & Rotating Reference Frame ........................ 36


3.1 Flow in Rotating Reference Frame .............................................................................................................. 36
3.1.1 Centrifugal & Coriolis Forces ............................................................................................................ 37
3.1.2 Relative Velocity Formulation .......................................................................................................... 37
3.1.3 Absolute Velocity Formulation......................................................................................................... 38
3.1.4 Early Formulation and Consideration ........................................................................................... 38
3.2 Flows with Rotating Reference Frames ..................................................................................................... 38
3.2.1 Single Rotating Reference Frame (SRF) Modeling ................................................................... 39
3.2.1.1 Case Study - Aerodynamics and Structural Analysis of (HAWT) Wind Turbine
Blade 40
3.2.1.1.1 Introduction .................................................................................................... 40
3.2.1.1.2 Airfoils and General Concepts of Aerodynamics ............................................ 41
3.2.1.1.3 Lift, Drag and Non-Dimensional Parameters .................................................. 41
3.2.1.1.4 Blade Element Momentum (BEM) Theory ..................................................... 43
3.2.1.1.5 Aerodynamic Load .......................................................................................... 43
3.2.1.1.6 Blade Geometry .............................................................................................. 43
3

3.2.1.1.7 Mathematic Model ......................................................................................... 44


3.2.1.1.8 Turbulence Model........................................................................................... 44
3.2.1.1.9 Mesh Generation ............................................................................................ 45
3.2.1.1.10 Result and Discussion ................................................................................... 46
3.2.1.1.11 Conclusion..................................................................................................... 47
3.2.2 Flow in Multiple Rotating Reference Frames (MRF) ............................................................... 47
3.2.2.1 Case Study – Mixing Tank ...................................................................................................... 48
3.2.3 The MRF Interface Formulation....................................................................................................... 49
3.2.3.1 Interface Treatment: Relative Velocity Formulation.................................................. 49
3.2.3.2 Interface Treatment: Absolute Velocity Formulation ................................................ 49
3.2.3.3 Case Study 1 - Experiments and CFD Calculations on the Performance of a
Non-Reversible Axial Fan ................................................................................................................................ 50
3.2.3.3.1 Introduction .................................................................................................... 50
3.2.3.3.2 Experimental Method ..................................................................................... 51
3.2.3.3.3 Numerical Method.......................................................................................... 51
3.2.3.3.4 Results and Discussion .................................................................................... 53
3.2.3.3.5 Conclusion....................................................................................................... 54
3.2.3.4 Case Study 2 - Potsdam Propeller CFD Benchmark .................................................... 55
3.2.3.4.1 Potsdam Propeller CFD Benchmark Description ............................................ 55
3.2.3.4.2 Meshing .......................................................................................................... 56
3.2.3.4.3 TCFD® Simulation Setup .................................................................................. 58
3.2.3.4.4 TCFD® Simulation Post-processing .................................................................. 58
3.2.3.4.5 Efficiency, Thrust & Torque Coefficient vs. Advance Coefficient ................... 59
3.2.3.4.6 Conclusion....................................................................................................... 59
3.2.3.4.7 References ...................................................................................................... 59
3.3 The Mixing Plane Model (MPM) .................................................................................................................... 60
3.3.1 Rotor and Stator Domains .................................................................................................................. 60
3.3.2 The Mixing Plane Concept .................................................................................................................. 61
3.3.3 Mixing Plane Algorithm ....................................................................................................................... 61
3.3.3.1 Mass Conservation Across the Mixing Plane.................................................................. 61
3.4 Sliding Mesh Modeling ...................................................................................................................................... 62
3.4.1 Sliding Mesh Theory ............................................................................................................................. 62
3.4.2 The Sliding Mesh Technique ............................................................................................................. 63
3.4.3 Sliding Mesh Concept ........................................................................................................................... 63

4 Elements of Turbomachinery ............................................................................................... 65


4.1 Background............................................................................................................................................................ 65
4.2 Historical Perspectives ..................................................................................................................................... 66
4.3 Modern Turbomachinery as Related to Gas Turbine Engine ............................................................ 66
4.4 Difference Among Turbojet, Turbofan and Turboprop Engines in Aviation .............................. 67
4.4.1 Turbojet ..................................................................................................................................................... 68
4.4.2 Turbofan .................................................................................................................................................... 68
4.4.3 Turboprop ................................................................................................................................................. 69
4.4.4 How does it work? ................................................................................................................................. 70
4.4.5 What is Thrust? ....................................................................................................................................... 71
4.5 Gas Turbine Performance ................................................................................................................................ 72
4.6 Gas Compressors ................................................................................................................................................. 73
4.6.1 Axial-Flow Compressors ..................................................................................................................... 73
4.6.2 Centrifugal Compressors .................................................................................................................... 73
4

4.7 Nomenclature of Terms .................................................................................................................................... 74


4.8 Component of Gas Turbine Engine .............................................................................................................. 77
4.8.1 Inlet .............................................................................................................................................................. 77
4.8.2 Axial Compressor ................................................................................................................................... 78
4.8.3 Diffuser ....................................................................................................................................................... 80
4.8.4 Nozzle ......................................................................................................................................................... 81
4.8.5 Combustor................................................................................................................................................. 81
4.8.6 Axial Gas Turbine ................................................................................................................................... 82
4.9 Difference in Blading Between Compressor and Turbine .................................................................. 83
4.10 Velocity Triangles in Turbomachines ......................................................................................................... 84
4.11 Energy Exchange with Moving Blades........................................................................................................ 84
4.11.1 Euler’s Equation for Turbomachinery .......................................................................................... 85
4.12 Compressors and their Reaction to Intake Distortion ......................................................................... 86
4.13 Effects of Turbine Temperature.................................................................................................................... 88
4.14 Compressor and Turbine Characteristics ................................................................................................. 90
4.14.1 Stall .............................................................................................................................................................. 90
4.14.2 Compressor Surge ................................................................................................................................. 91
4.14.3 Choked Flow............................................................................................................................................. 91
4.15 Additional Types of Turbines ......................................................................................................................... 92

5 Primary Research in Turbomachinery.............................................................................. 93


5.1 Research Spectrum ............................................................................................................................................. 93
5.2 Application of CFD in Turbomachinery ..................................................................................................... 94
5.3 Quasi 3D Flow (Q3D) ......................................................................................................................................... 94
5.3.1 Stream Surface of Second Kind - Through Flow (S2) .............................................................. 95
5.3.2 Stream Surface of First Kind (Blade 2 Blade – S1) ................................................................... 96
5.3.3 Case Study – Turbine Airfoil Optimization Using Inviscid Quasi 3D (Q3D) Analysis
Codes 97
5.3.3.1 Quasi-3D CFD Analysis and Results ................................................................................... 98
5.4 Theory of Radial Equilibrium in Through Flow (Cr = 0) .................................................................. 100
5.5 Governing Equation of Rotating Frame of Reference ....................................................................... 101
5.6 Efficiency Effects in Turbomachinery...................................................................................................... 102
5.6.1 Isentropic Efficiency .......................................................................................................................... 103
5.7 Case Study 1 – Computation of Heat Transfer in Linear Turbine Cascade............................... 104
5.7.1 Numerical Methods ............................................................................................................................ 105
5.7.2 Mesh Generation ................................................................................................................................. 105
5.7.3 Heat Transfer Results for 2D & 3D .............................................................................................. 106
5.7.4 Experimental Data .............................................................................................................................. 107
5.7.5 Effects of Turbulence......................................................................................................................... 108
5.8 Case Study 2 - Using Shock Control Bumps To Improve Transonic Compressor Blade
Performance ................................................................................................................................................................... 109
5.8.1 Introduction and Motivation .......................................................................................................... 110
5.8.2 Shock Control for Turbomachinery & Literature Survey ................................................... 110
5.8.3 Shock Control Bumps ........................................................................................................................ 111
5.8.4 Test Case - NASA Rotor 37 .............................................................................................................. 112
5.8.5 Validation ............................................................................................................................................... 113
5.8.6 Flow Field For the Datum Case ..................................................................................................... 114
5.8.7 Validation ............................................................................................................................................... 115
5.8.8 Blade Flow Features .......................................................................................................................... 115
5

5.8.9 Adjoint Sensitivity Analysis ............................................................................................................ 116


5.8.10 Shock Bump Parameterization & Optimization ..................................................................... 117
5.8.11 Optimization Method......................................................................................................................... 118
5.8.12 Rotor 37 Bump Optimization ......................................................................................................... 118
5.8.13 Analysis of the R37 Optimized Bump Design .......................................................................... 119
5.8.14 Performance Across the Characteristic for R37 ..................................................................... 120
5.8.15 Conclusion.............................................................................................................................................. 121

6 Complex Flow in Turbomachinery................................................................................... 123


6.1 Key Features of Transonic Fan (Turbine) Field .................................................................................. 123
6.2 Sources of Unsteadiness in Turbomachinery ....................................................................................... 124
6.3 Interaction of Potential Flows in Adjacent Blade Rows ................................................................... 126
6.3.1 Interactions in Transonic Fan ........................................................................................................ 126
6.4 Interaction Between Wake Flow and Blade Rows.............................................................................. 127
6.5 Interaction Between Secondary Flows and Blade Rows.................................................................. 127
6.6 Wake-Boundary Layer Interaction ........................................................................................................... 128
6.7 Un-Shrouded Tip Leakage Flow Interaction ......................................................................................... 129
6.8 General Review on Secondary Flows ....................................................................................................... 130
6.8.1 Classical View ....................................................................................................................................... 131
6.8.2 Modern View ......................................................................................................................................... 132
6.8.3 Latest View ............................................................................................................................................ 134
6.8.4 Comparing and Contrasting Secondry Flow in Turbine and Compressors................. 135
6.9 3D Separation .................................................................................................................................................... 136
6.9.1 Compressors Example ...................................................................................................................... 137
6.10 Airfoil End-Wall Heat Transfer ................................................................................................................... 138
6.10.1 Theoretical Development of End-Wall Flows.......................................................................... 139
6.10.2 End-Wall Heat Transfer.................................................................................................................... 141
6.10.3 Leading Edge Modifications............................................................................................................ 142
6.10.4 Blade Tip Heat Transfer ................................................................................................................... 143
6.10.5 Case Study 1 - Effects of Grid Refinement and Turbulence in 3D Flow Structure and
End-Wall Heat Transfer in Transonic Turbine Blade Cascade ............................................................ 144
6.10.5.1 Problem Definition ................................................................................................................ 144
6.10.5.2 Computational Aspects ........................................................................................................ 145
6.10.5.3 Results and Discussion......................................................................................................... 146
6.10.5.4 End-Wall Heat Transfer Sensitivity w.r.t Grid and Turbulence Models .......... 147
6.10.5.5 Summary .................................................................................................................................... 149
6.10.6 Case Study 2 - Comparison of Steady and Unsteady RANS Heat Transfer Simulations
of Hub and End all of a Turbine Blade Passage ......................................................................................... 150
6.10.6.1 Introduction.............................................................................................................................. 150
6.10.6.2 Computational Method ........................................................................................................ 152
6.10.6.2.1 Boundary Conditions .................................................................................. 153
6.10.6.3 Results and Discussion......................................................................................................... 154
6.10.6.4 Heat Transfer Comparison With Open Literature .................................................... 162
6.10.6.4.1 Comparison of Casing Heat Flux With [Epstein et al.] ................................ 163
6.10.6.4.2 Comparison of Hub Heat Transfer With Tallman et al. .............................. 163
6.10.6.5 Conclusions ............................................................................................................................... 164

7 Blade Cooling ........................................................................................................................... 166


7.1 Film Cooling Effects......................................................................................................................................... 167
6

7.1.1 Fundamentals of Film Cooling ....................................................................................................... 167


7.1.1.1 End-Wall Film-Cooling ......................................................................................................... 169
7.2 Blade Cooling Using Vanes in Blades ....................................................................................................... 171
7.2.1 Vane NASA C3X .................................................................................................................................... 172
7.2.2 Showered Film Type Cooling ......................................................................................................... 173
7.2.3 Complex Network of Vanes ............................................................................................................. 173
7.3 Conjugate Heat Transfer ............................................................................................................................... 174
7.3.1 Case Study - Heat Transfer in Separated Flows on the P. S. of Turbine Blades ........ 174
7.3.1.1 Literature Survey ................................................................................................................... 174
7.3.1.2 CFD Modeling ........................................................................................................................... 175
7.3.1.3 Description of the Blade Computational Grids and Results for Attached Flow
176
7.3.1.4 Separated Flow with Large Separation Bubble ........................................................ 177
7.3.1.5 Inlet Flow Angle Effects ....................................................................................................... 180
7.3.1.5.1 Reynolds Number Effect ............................................................................... 181
7.3.1.6 Concluding Remarks ............................................................................................................. 183

8 General Perspectives on Turbulence Consideration ................................................. 185


8.1 Case Study 1 - Turbulence Comparisons for a Low Pressure 1.5 Stage Test Turbine ........ 187
8.2 Case Study 2 - Comparison of Various Turbulence Models in Rotating Machinery Blade-to-
Blade Passages............................................................................................................................................................... 187
8.2.1 Introduction .......................................................................................................................................... 187
8.2.2 Turbulent Models................................................................................................................................ 188
8.2.2.1 Baldwin-Lomax’s Zero-Equation Model [1] ................................................................ 188
8.2.2.2 Birch’s One-Equation Model [3] ....................................................................................... 188
8.2.2.3 Standard Two-Equation Model [5] ................................................................................. 188
8.2.3 Numerical Results ............................................................................................................................... 189
8.2.3.1 Transonic Compressor ......................................................................................................... 189
8.2.3.2 UTRS Turbine Blades ............................................................................................................ 189
8.2.3.3 C4 Compressor ........................................................................................................................ 190
8.2.4 Concluding Remarks .......................................................................................................................... 191
8.2.5 References.............................................................................................................................................. 191

9 Rotor-Stator Interaction Treatment (RST) ................................................................... 193


9.1 Physical Perspectives ..................................................................................................................................... 193
9.2 Multi-Passage vs. Multi-Stages ................................................................................................................... 194
9.3 Case for Mixing Plane Model ....................................................................................................................... 195
9.4 Steady Treatment of Interface (Mixing Plane) ..................................................................................... 196
9.4.1 Losses Across the Interface of Mixing Plane ............................................................................ 197
9.4.2 Principles of Flux Conservation .................................................................................................... 198
9.4.3 Case Study 1 - Comparison of Flux Balanced Mixing Models on Q-1.5 Stage Rotor 67
200
9.4.4 Case Study 2 - Modeling of Secondary Flows in Single Blade Rows using Mixing
Plane Approach ....................................................................................................................................................... 202
9.4.4.1 Transonic Turbine Stage Meshing and Flow Details ............................................... 203
9.4.5 Case Study 3 - Improvement Methods for Mixing Plane Models ..................................... 204
9.4.5.1 Validation Test Case .............................................................................................................. 205
9.4.5.2 1.5 Stage Transonic Axial Compressor .......................................................................... 205
9.4.6 Frozen Rotor ......................................................................................................................................... 206
7

9.5 Un-Steady Treatment of Interface............................................................................................................. 207


9.5.1 Sliding Mesh (MRF) ............................................................................................................................ 207
9.5.2 Non-Linear Harmonic Balanced Method (NLHB) .................................................................. 208
9.5.3 Profile Transformation (Pitch Scaling) ...................................................................................... 209
9.5.4 Time Transformation Method (TT) using Phase-Shifted Periodic Boundary
Conditions 210
9.5.5 Revisiting Non-Linear Harmonic Balance (NLHB) Methodology.................................... 212
9.5.5.1 Temporal & Spatial Periodicity Requirement ............................................................ 212
9.5.5.2 Boundary Conditions ............................................................................................................ 212
9.5.5.3 Fourier 'Shape Correction' for Single Passage Time-Marching Solution ........ 215
9.5.5.4 Case Study 1 – 2D Compressor Stage ............................................................................. 216
9.5.5.5 Case Study 2 - 3D Flow in Turbine Cascade ................................................................ 216
9.5.6 Assessment of 2D Steady and Unsteady Adjoint Sensitivities for Rotor-Starter
Interaction 217
9.5.6.1 Case Study ................................................................................................................................. 218
9.5.6.2 Results......................................................................................................................................... 219
9.5.6.2.1 Case 1 - Subsonic Stage ................................................................................ 220
9.5.6.2.2 Case 2 - Transonic Stage ............................................................................... 220
9.5.6.3 Design Sensitivity ................................................................................................................... 221
9.5.6.3.1 Case1 - Subsonic Stage ................................................................................. 222
9.5.6.3.2 Case2 - Transonic Stage ................................................................................ 223
9.5.6.4 Conclusions ............................................................................................................................... 223
9.6 Case Study - Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design
Perspective...................................................................................................................................................................... 224
9.6.1 Introduction .......................................................................................................................................... 224
9.6.2 Stator-Rotor Interaction in Axial Stages .................................................................................... 226
9.6.2.1 Stator Wake-Rotor Blade Interaction .......................................................................... 226
9.6.2.2 Stator Secondary Lows-Rotor Blade Interaction ...................................................... 227
9.6.2.3 Off-Design Conditions ........................................................................................................... 234
9.6.2.4 Stator Shock-Rotor Blade Interaction ............................................................................ 235
9.6.3 Design Perspective ............................................................................................................................. 237
9.6.3.1 Axial Gap .................................................................................................................................... 237
9.6.3.2 End wall Contouring and 3D Blade Geometries ........................................................ 238
9.6.3.3 Cascades Clocking .................................................................................................................. 239

10 Axial Turbo-Machinery Design and Optimization ...................................................... 241


10.1 A Road Map to Turbo-Machine Design and Optimization ............................................................... 241
10.2 Optimization Methods for Turbomachinery Designs........................................................................ 243
10.2.1 Wu’s Pioneering (S1 and S2) Scheme ......................................................................................... 244
10.2.2 Concept of Streamline Curvature Method ................................................................................ 244
10.3 Case Study 1 - Aerodynamic Design of Compressors ........................................................................ 245
10.3.1 Statement of Problem........................................................................................................................ 245
10.3.2 Different Compressors Objectives ............................................................................................... 246
10.3.3 Design Techniques for Compressor ............................................................................................ 247
10.3.4 Preliminary Design Techniques (1D) ......................................................................................... 249
10.3.5 Through Flow Design Techniques (2D) ..................................................................................... 249
10.3.6 Detailed Design Techniques (3D)................................................................................................. 250
10.3.6.1 Direct Methods ........................................................................................................................ 250
10.3.6.2 Inverse Methods ..................................................................................................................... 251
8

10.3.7 Concluding Remarks .......................................................................................................................... 252


10.4 Case Study 2 – Turbine Airfoil Optimization using Quasi 3D Analysis Codes ......................... 253
10.4.1 Parametric Representation of Airfoil Design Process ......................................................... 254
10.4.2 Constraints and Problem Formulation ...................................................................................... 255
10.4.3 Quasi-3D CFD Analysis and Results ............................................................................................ 257
10.4.4 Concluding Remarks .......................................................................................................................... 258
10.5 Case Study 3 – 2D Design Optimization of Turbine Blade in Quasi-Periodic Unsteady Flow
Problems Using a Harmonic Balance Method .................................................................................................. 259
10.5.1 OptC1 Configuration .......................................................................................................................... 260
10.5.2 Optimization Problem and Results .............................................................................................. 261
10.6 Case Study 4 - Using Shock Control Bumps To Improve Transonic Compressor Blade
Performance ................................................................................................................................................................... 263
10.6.1 Introduction and Motivation .......................................................................................................... 264
10.6.2 Shock Control for Turbomachinery & Literature Survey ................................................... 264
10.6.3 Shock Control Bumps ........................................................................................................................ 265
10.6.4 Test Case - NASA Rotor 37 .............................................................................................................. 266
10.6.5 Validation ............................................................................................................................................... 267
10.6.6 Flow Field For the Datum Case ..................................................................................................... 268
10.6.7 Validation ............................................................................................................................................... 269
10.6.8 Blade Flow Features .......................................................................................................................... 269
10.6.9 Adjoint Sensitivity Analysis ............................................................................................................ 270
10.6.10 Shock Bump Parameterization & Optimization ..................................................................... 271
10.6.11 Optimization Method......................................................................................................................... 272
10.6.12 Rotor 37 Bump Optimization ......................................................................................................... 272
10.6.13 Analysis of the R37 Optimized Bump Design .......................................................................... 273
10.6.14 Performance Across the Characteristic for R37 ..................................................................... 274
10.6.15 Conclusion.............................................................................................................................................. 275
10.7 Case Study 5 - 3D Multi-Objective Design Optimization of a Transonic Compressor Rotor
276
10.7.1 Introduction .......................................................................................................................................... 276
10.7.2 Influence of Blade Shape on Transonic Compressor Flows .............................................. 277
10.7.2.1 Effect of Blade Profiles ......................................................................................................... 277
10.7.2.2 Effect of 3-D Sweep and Lean ............................................................................................ 278
10.7.3 Optimization of NASA Rotor 37 .................................................................................................... 279
10.7.3.1 Blade Geometry Definition ................................................................................................. 280
10.7.3.2 Flow Solver: Description and Validation ...................................................................... 281
10.7.3.3 Multi-Objective Evolutionary Algorithm ...................................................................... 283
10.7.4 Results ..................................................................................................................................................... 284
10.7.5 Conclusions............................................................................................................................................ 286
10.8 Case Study 6 - Effect of Twist of a Wide-Chord Fan-Blade on the Aerodynamic Performance
of the Fan of a High-Bypass Turbofan Engine .................................................................................................. 287
10.8.1 Introduction .......................................................................................................................................... 287
10.8.2 Numerical Models and Technique ............................................................................................... 289
10.8.2.1 Model Construction ............................................................................................................... 289
10.8.2.2 Domain Construction and Meshing................................................................................. 289
10.8.2.3 Numerical Scheme ................................................................................................................. 289
10.8.2.4 Methodology and Time Stepping ..................................................................................... 291
10.8.3 Numerical Results ............................................................................................................................... 291
10.8.3.1 Performance of All Cases ..................................................................................................... 291
10.8.3.2 Blade Twist Effect at High Mass Flow Rate Conditions .......................................... 292
9

10.8.3.3 Blade Twist Effect at Low Mass Flow Rate Conditions ........................................... 295
10.8.4 Conclusion.............................................................................................................................................. 296
10.8.5 References.............................................................................................................................................. 297

11 Radial Flow ............................................................................................................................... 299


11.1 Centrifugal Compressor................................................................................................................................. 299
11.1.1 Operation Theory................................................................................................................................ 299
11.1.2 Similarities to Axial Compressor .................................................................................................. 299
11.1.3 Components of a simple Centrifugal Compressor ................................................................. 300
11.1.3.1 Inlet .............................................................................................................................................. 300
11.1.3.2 Centrifugal Impeller .............................................................................................................. 300
11.1.3.3 Diffuser ....................................................................................................................................... 301
11.1.3.4 Collector ..................................................................................................................................... 301
11.1.4 Applications........................................................................................................................................... 302
11.1.4.1 Gas Turbines and Auxiliary Power Units ...................................................................... 302
11.1.4.2 Automotive and Diesel Engines Turbochargers and Superchargers ................ 302
11.1.4.3 Natural Gas to Move the Gas from the Production site to the Consumer ....... 302
11.1.4.4 Oil Refineries, Natural Gas Processing, Petrochemical and Chemical Plants 302
11.1.4.5 Air-Conditioning and Refrigeration and HVAC .......................................................... 302
11.1.4.6 Industry and Manufacturing to Supply Compressed Air ....................................... 302
11.1.4.7 Air Separation Plants to Manufacture Purified End Product Gases .................. 303
11.1.4.8 Oil Field Re-Injection of High Pressure Natural Gas to Improve Oil Recovery
303
11.2 Radial Turbine ................................................................................................................................................... 303
11.2.1 Advantages and Challenges ............................................................................................................ 303
11.2.2 Types of Radial Turbines ................................................................................................................. 304
11.2.2.1 Cantilever Radial Turbine ................................................................................................... 304
11.2.2.2 90 Degree IFR Turbine ......................................................................................................... 304
11.2.2.3 Outward-Flow Radial Stages ............................................................................................. 304

12 Best Procedures for Turbomachinery ............................................................................ 306


12.1 Quasi-3D (Q3D) or 3D Simulation ............................................................................................................. 306
12.1.1 2D Simulations ..................................................................................................................................... 306
12.1.2 Quasi-3D (Q3D) Simulation ............................................................................................................ 306
12.1.3 Full 3D Simulations ............................................................................................................................ 306
12.2 Single vs Multi-Stage Analysis .................................................................................................................... 307
12.2.1 Single Stage ............................................................................................................................................ 307
12.2.2 Multi-Stage Analysis .......................................................................................................................... 308
12.2.2.1 Steady Mixing-Plane Simulations .................................................................................... 308
12.2.2.2 Steady Frozen Rotor Simulations .................................................................................... 308
12.2.2.3 Unsteady Sliding Mesh Stator-Rotor Simulations ..................................................... 308
12.2.2.4 Unsteady Harmonic Balance Simulations .................................................................... 308
12.2.2.5 Hybrid Steady-Unsteady Stator-Rotor Simulations ................................................. 309
12.2.2.6 Other Advanced Multi-Stage Methods ........................................................................... 309
12.3 Inviscid or Viscid .............................................................................................................................................. 309
12.4 Transient or Steady-State ............................................................................................................................. 309
12.5 Meshing ................................................................................................................................................................ 311
12.5.1 Mesh Size Guidelines ......................................................................................................................... 312
12.5.2 Case Study - Mesh Resolution Effect on 3D RANS Turbomachinery Flow Simulations
10

313
12.5.2.1 Formulation of Problems .................................................................................................... 313
12.5.2.2 Conclusions ............................................................................................................................... 314
12.5.3 Boundary Mesh Resolution ............................................................................................................. 314
12.5.4 Periodic Meshing ................................................................................................................................. 315
12.6 Boundary Conditions ...................................................................................................................................... 315
12.7 Turbulence Modeling...................................................................................................................................... 317
12.8 Aero-Mechanics ................................................................................................................................................ 317
12.8.1 Nodal Diameter .................................................................................................................................... 319
12.9 Near Wall Treatment ...................................................................................................................................... 319
12.10 Transition Prediction ..................................................................................................................................... 319
12.11 Numerical Consideration .............................................................................................................................. 320
12.12 Convergence Criteria ...................................................................................................................................... 320
12.13 Single or Double Precision ........................................................................................................................... 320
12.14 Heat Transfer Prediction .............................................................................................................................. 320
12.14.1 Keeping it Cool in Gas Turbine ...................................................................................................... 321
12.15 Literature Review and Parallel Processing Tools ............................................................................... 321
12.16 Concluding Remarks ....................................................................................................................................... 322

List of Tables

Table 3.1 Mesh Resolution and Number of Cells ........................................................................................... 52


Table 3.2 Solution Methods and Boundary Conditions ............................................................................... 53
Table 3.3 Prescribed Boundary zone for Mixing Plane ............................................................................... 61
Table 4.1 Glossary of Turbomachinery Terms ............................................................................................... 74
Table 5.1 Rotor 37 Optimized Bump Performance Comparison – Courtesy of [John et al.] ..... 119
Table 6.1 Parameters of the Grids Used ......................................................................................................... 145
Table 9.1 Rotor/Stator Interaction Schemes ................................................................................................ 194
Table 9.2 Nomenclature for different Mixing Models Used in Study.................................................. 200
Table 9.3 Axial turbine simulation parameters ........................................................................................... 218
Table 10.1 Axial Flow Compressor Design .................................................................................................... 246
Table 10.2 Airfoil Geometry Parameters ....................................................................................................... 255
Table 10.3 Airfoil Design Variables .................................................................................................................. 256
Table 10.4 Rotor 37 Optimized Bump Performance Comparison – Courtesy of [John et al.] .. 273
Table 10.5 Approximated model dimensions and specifications ........................................................ 289

List of Figures

Figure 1.1 Layout of the KJ66 micro gas turbine ........................................................................................... 21


Figure 1.2 Three Types of Full Engine Simulation Methodologies ......................................................... 22
Figure 2.1 Vortex created by the passage of an aircraft wing, revealed by colored smoke ......... 23
Figure 2.2 Rigid-Body Vortex ................................................................................................................................. 24
Figure 2.3 3D Visualization of a Vortex Curve ................................................................................................ 25
Figure 2.4 A Plughole Vortex .................................................................................................................................. 25
Figure 2.5 Types of Impeller .................................................................................................................................. 26
Figure 2.6 Flow Direction of Three Different Pumps/Impellers (Courtesy of Global Spec) ....... 27
Figure 2.7 A centrifugal pump uses an impeller with backward-swept arms ................................... 27
Figure 2.8 Axial Flow Impeller (left) and Radial Flow Impeller (right) ............................................... 27
Figure 2.9 Centrifugal Pumps ................................................................................................................................ 29
Figure 2.10 Sketch of the experimental set-up ............................................................................................... 30
11

Figure 2.11 For s ≥ 0 co-rotation at different speed ..................................................................................... 32


Figure 2.12 For s < 0 counter-rotating at different speed......................................................................... 32
Figure 2.13 Geometry of Swept and Dihedral Blades .................................................................................. 33
Figure 2.14 Experimental Pressure Iso-Surfaces ; Left - Without Sweep ;Right - With Forward
Sweep (Courtesy of RÁBAI and VAD [93])............................................................................................................ 34
Figure 2.15 Linear Turbine Cascade ................................................................................................................... 34
Figure 3.1 Rotating Frame of Reference ............................................................................................................ 36
Figure 3.2 Centrifugal and Coriolis Force ......................................................................................................... 37
Figure 3.3 Single Blade Model with Rotationally Periodic Boundaries ................................................ 39
Figure 3.4 Airfoil Cross-Sections used in the Design of the Wind Turbine Blades .......................... 41
Figure 3.5 Typical Lift and Drag Coefficients .................................................................................................. 42
Figure 3.6 (Top) Airfoils Superposed on the Wind Turbine Blade and (Bottom) Top View of a
Subset of the Airfoil Cross-Sections illustrating Blade Twisting ................................................................. 44
Figure 3.7 Force Analysis for S818 Airfoil Section ........................................................................................ 45
Figure 3.8 Static Pressure, Velocity Magnitudes, Deformation and Stress Distribution ............... 46
Figure 3.9 Mixing Tank Geometry with One Rotating Impeller ............................................................... 48
Figure 3.10 Mixing Tank with Two Rotating Impellers .............................................................................. 48
Figure 3.11 Interface Treatment for the MRF Model ................................................................................... 49
Figure 3.12 An Industrial Axial Fan Used For Ventilation ......................................................................... 50
Figure 3.13 AMCA 210 test rig [8]........................................................................................................................ 51
Figure 3.14 Surface mesh of the rotating zone ............................................................................................... 52
Figure 3.15 Cell zones and boundary conditions ........................................................................................... 53
Figure 3.16 The streamlines inside the duct at a rotor speed of 1500 rpm ....................................... 54
Figure 3.17 Torque vs. Volumetric Flow Rate Curves ................................................................................. 54
Figure 3.18 SVA Potsdam Laboratory ................................................................................................................ 55
Figure 3.19 Potsdam Propeller Benchmark - POINTWISE Mesh ..................................................... 56
Figure 3.20 Domain Decomposition of MRF .................................................................................................... 57
Figure 3.21 Potsdam Propeller Benchmark - snappyHexMesh Mesh........................................... 57
Figure 3.22 Surface Static Pressure ..................................................................................................................... 58
Figure 3.23 Comparison of TCFD vs. AVA using Pointwise Mesh ........................................................... 59
Figure 3.25 Mixing Plane Concepts as Applied to Axial Rotation ........................................................... 60
Figure 3.24 Mixing Plane Concepts Applied to Radial Rotation .............................................................. 61
Figure 3.26 Illustration of Unsteady Interactions .......................................................................................... 62
Figure 3.27 Examples of Transient Interaction using Sliding Mesh ...................................................... 63
Figure 3.28 Initial position and some translation with Sliding Interface ............................................ 64
Figure 4.1 Classification of Turbomachines ..................................................................................................... 65
Figure 4.2 Component of Turbomachines and their Thermodynamic (Brayton cycle) properties
67
Figure 4.3 Turbojet Engine ..................................................................................................................................... 68
Figure 4.4 Turbofan Engine .................................................................................................................................... 68
Figure 4.5 Turboprop Engine................................................................................................................................. 69
Figure 4.6 Turboshaft Engine ................................................................................................................................ 69
Figure 4.7 Twin Pool Trubofan Jet Engine........................................................................................................ 70
Figure 4.8 A 1D Control Volume around a propulsion system (Courtesy’s of NASA Glen
Research Center) ............................................................................................................................................................. 71
Figure 4.9 Gas Compressor Types........................................................................................................................ 72
Figure 4.10 A single stage Centrifugal Compressor ...................................................................................... 74
Figure 4.11 Schematics of Axial Compressor .................................................................................................. 73
Figure 4.12 Blade Related Terminology ............................................................................................................ 77
Figure 4.13 Schematic Diagram of fluid properties through an axial compressor stage –
12

Courtesy of [T. B. Ferguson, Gravdahl, and Egeland] ..................................................................................... 78


Figure 4.14 Pressure and Velocity profile through a Multi-Stage Axial Compressor ..................... 79
Figure 4.15 Combustor Primary Operating Components .......................................................................... 81
Figure 4.16 Schematics of Axial Flow Turbine ............................................................................................... 82
Figure 4.17 Turbine Flow Characteristics ........................................................................................................ 82
Figure 4.18 Examples of Typical Blades for Compressor and Turbine ................................................ 83
Figure 4.19 Velocity triangles for an Axial Compressor ............................................................................. 84
Figure 4.20 Velocity Triangles in Relation to Incident Angle ................................................................... 86
Figure 4.21 Compressor Operating Map ........................................................................................................... 87
Figure 4.22 Sample engine Perssure, Velocity and Temperature variation ....................................... 88
Figure 4.23 Turbine Inlet Temperature27 ......................................................................................................... 89
Figure 4.24 Characteristics Graph of a Compressor ..................................................................................... 90
Figure 4.25 Classical Compressor Surge Cycles ............................................................................................. 91
Figure 4.26 Illustration of the Propagation of a Stall Cell in the Relative Frame ............................. 91
Figure 5.1 Impact of CFD on SNECMA fan performance, over a period of 30 years........................ 93
Figure 5.2 Illustration of S1 and S2 surfaces ................................................................................................... 95
Figure 5.3 Streamline Curvature Method ......................................................................................................... 96
Figure 5.4 The turbine Design Process .............................................................................................................. 98
Figure 5.5 Flow Path of the Turbine.................................................................................................................... 99
Figure 5.6 Schematics of an airfoil showing stream lines along the radial direction .................. 100
Figure 5.7 3D model of an airfoil showing the passage between adjacent airfoils ....................... 100
Figure 5.8 Radial Equilibrium ............................................................................................................................. 100
Figure 5.9 Coriolis and Centripetal forces created by the Rotating Frame of Reference ........... 101
Figure 5.10 Compression process ..................................................................................................................... 103
Figure 5.11 Expansion process........................................................................................................................... 103
Figure 5.12 Linear Turbine Cascade and Computational Domain....................................................... 104
Figure 5.13 A) Default mesh B) Refine mesh ....................... 105
Figure 5.14 Average Pressure Specification at pressure boundary .................................................... 106
Figure 5.15 Stanton Number Distribution on End-Wall .......................................................................... 107
Figure 5.16 Stanton number Distribution on Blade Surface for 2D Grid .......................................... 108
Figure 5.17 Contours Of Casing Static Pressure Beneath A High-Speed Rotor (550 M/S ......... 109
Figure 5.18 Schematic of Shock Structures (A) Datum, (B) S-Shaped Design. Courtesy of [John
et al.] 111
Figure 5.19 Datum Geometry and Optimized Shock Control Bumps on The Mid- Section of
Nasa Rotor 67- From Mazaheri et al.. ................................................................................................................. 112
Figure 5.20 The R37 CFD Domain Used – Courtesy of [John et al.] ..................................................... 113
Figure 5.21 Radial Profiles Vs Experimental Data – Courtesy of [John et al.] ................................. 114
Figure 5.22 (a) 3d Separation (Orange) On The R37 Geometry (Flow Right To Left), (b) Rel.
Mach No. Contour At 60% Span – Courtesy of [John et al.]......................................................................... 115
Figure 5.23 Shock Region Flow Features For RR-Fan At Points A) A, B) B, C) C, D) D, E) E, F) F.
Flow Direction – Courtesy of [John et al.] .......................................................................................................... 116
Figure 5.24 Example 2d CST Bump (Solid Line) And The Four Polynomials Used To Construct
It (Dashed Lines) – Courtesy of [John et al.]...................................................................................................... 117
Figure 5.25 a) Example Individual Bump Geometry ................................................................................. 117
Figure 5.26 Spanwise Slice Of The Datum And Optimized R37 Geometries At 60% Span –
Courtesy of [John et al.].............................................................................................................................................. 118
Figure 5.27 Optimized R37 Bump (Blue) Added To The Datum Blade Geometry (Grey) ......... 118
Figure 5.28 Datum (Left) And Optimized (Right) Rotor 37 Static Pressure Contours. Flow
Direction Right To Left – Courtesy of [John et al.] .......................................................................................... 119
Figure 5.29 Datum (Left) And R37 Optimized (Right) ............................................................................. 120
13

Figure 5.30 Datum (Left) And Optimized (Right) Rotor 37 Separated Flow Contours (Orange).
Flow Direction Right To Left.................................................................................................................................... 120
Figure 5.31 Lift Plots For The Datum and Optimized ............................................................................... 121
Figure 5.32 R37 Optimized Characteristic Vs Datum – Courtesy of [John et al.] ........................... 121
Figure 6.1 Complex Flow Phenomena Compressors ................................................................................. 123
Figure 6.2 Fan Tip Section Geometry .............................................................................................................. 124
Figure 6.3 Flow Structures with 5 to 6 Orders of Magnitudes Variations in Length and Time
Scales (LaGraff et al., 2006)...................................................................................................................................... 125
Figure 6.4 Shock Structure in Transonic Fan ............................................................................................... 126
Figure 6.5 Pressure Contour of Wake Flow .................................................................................................. 127
Figure 6.6 Instantaneous Absolute Velocity Contour at Nozzle Exit [Matsunuma, 2006]......... 128
Figure 6.7 Unsteady Wakes Convecting in Blade Passage ...................................................................... 129
Figure 6.8 Flow over an unshrouded tip gap ............................................................................................... 130
Figure 6.9 Classical Secondary Flow Model .................................................................................................. 131
Figure 6.10 Modern Secondary Flow Model ................................................................................................. 132
Figure 6.11 Vortex pattern of Latest Secondary Flows ............................................................................ 133
Figure 6.12 Turbine Secondary Flow Model (Takeishi et al.) ............................................................... 135
Figure 6.13 Illustration of formation of hub corner stall together with limiting streamlines and
separation lines ............................................................................................................................................................. 137
Figure 6.14 Three-dimensional separations: traditional view and scope of current
investigation ................................................................................................................................................................... 137
Figure 6.15 Illustration of the near wall flows as taken through oil and dye surface flow
visualization (reproduced with permission of the publisher from ASME) ......................................... 138
Figure 6.16 (Top) - Measurements of the Horseshoe Vortex just upstream of the Vane at the
Vane-End-Wall Juncture (Bottom) - Actual Hardware Showing Effects of the Horseshoe Vortex
on a First Vane (Courtesy of ASME) .................................................................................................................... 139
Figure 6.17 Illustration of Different Vortical ................................................................................................ 140
Figure 6.18 Contours of Non-Dimensional Heat Transfer Coefficients (Reproduced with
Permission ASME) ...................................................................................................................................................... 141
Figure 6.19 Fillet and Bulb Designs as Shown by (Becz et al.) ............................................................... 142
Figure 6.20 CFD Prediction of Streamlines Across a ................................................................................. 143
Figure 6.21 Blade Passage and Slice of The Computational Domain.................................................. 144
Figure 6.22 Computed Mid-Span Mach Number Distribution............................................................... 146
Figure 6.23 End-Wall Stanton Number (103) Distributions Computed with Grid B in
Comparison with the Measurement Data: (1) k-ω Turbulence Model, (2) M-SST, (3) v2-f, (4)
Experiment ..................................................................................................................................................................... 147
Figure 6.24 Effect of Grid Refinement on the End-Wall Stanton Number (x103) Prediction with
the M-SST Turbulence Model: (1) Grid С, (2) Grid D, (3) Grid E, (4) Experiment. ............................ 148
Figure 6.25 End-Wall Streak Line Visualization.......................................................................................... 149
Figure 6.26 Grid on the Solid Surfaces of the Geometry .......................................................................... 152
Figure 6.27 (a) Casing surface mesh showing multiblock structure and (b) hub surface mesh
showing multiblock structure ................................................................................................................................. 153
Figure 6.28 Total temperature (T0) and total pressure (P0) at the blade inlet ............................... 154
Figure 6.29 Instantaneous and time-averaged „dashed… hub surface.............................................. 156
Figure 6.30 Instantaneous and time-averaged (dashed) hub heat transfer distribution for a
wake passing .................................................................................................................................................................. 155
Figure 6.31 Time-Averaged Casing Pressure ............................................................................................... 157
Figure 6.32 Time-Averaged Hub Pressure .................................................................................................... 157
Figure 6.33 Difference in hub pressure distribution between the time-averaged and steady
results................................................................................................................................................................................ 158
14

Figure 6.34 Difference in casing pressure distribution between the time-averaged and steady
results................................................................................................................................................................................ 158
Figure 6.35 Time-averaged casing adiabatic wall temperature ........................................................... 159
Figure 6.36 Time-averaged hub adiabatic wall temperature ................................................................ 159
Figure 6.37 Difference in casing adiabatic wall temperature distribution between the time-
averaged and steady results .................................................................................................................................... 160
Figure 6.38 Difference in hub adiabatic wall temperature distribution between the time-
averaged and steady results .................................................................................................................................... 160
Figure 6.39 Time-averaged casing heat transfer rate ............................................................................... 161
Figure 6.40 Time-averaged hub heat transfer rate .................................................................................... 161
Figure 6.41 Difference in casing heat transfer rate between the time-averaged and steady
results................................................................................................................................................................................ 162
Figure 6.42 Difference in hub heat transfer rate between the time averaged and steady results
162
Figure 6.43 Difference in hub Nusselt number between the time averaged and steady results
163
Figure 7.1 The Schematic of a Modern Gas Turbine Blade with Common Cooling Techniques
(Courtesy of Je-Chin Han) .......................................................................................................................................... 166
Figure 7.2 Typical high-pressure turbine stage showing rim seal and wheel-space ................... 167
Figure 7.3 Schematic of Flm Cooling Concept .............................................................................................. 168
Figure 7.4 Measured Adiabatic Wall Temperatures for Coolant Exiting a Combustor/Vane
Leakage Slot (reproduced with permission from ASME) ............................................................................ 169
Figure 7.5 Contours of Adiabatic Effectiveness for Two Film-Cooling Hole Patterns (left and
center) With a Mid-Passage Gutter for the Cooling Hole Pattern in the Center (Right)
(Reproduced With Permission From the Publisher of ASME) .................................................................. 170
Figure 7.6 Vane Section with Ten Cooling Channels and Temperature Distribution Computed
171
Figure 7.7 Surface Temperature Distribution on the Suction Side (Left) and The Pressure Side
(Right) of the Vane ....................................................................................................................................................... 172
Figure 7.8 Flow Streamlines Colored Cooling Air Temperature into Passages ............................. 173
Figure 7.9 2D Hybrid Mesh around the T106 Blade ................................................................................. 177
Figure 7.10 Blade Profile vs. Pressure Coefficient (Courtesy of De La Calzada et al.) ................. 177
Figure 7.11 Flow Field at the Front and Middle Parts of the Separation Bubble (Courtesy of De
La Calzada et al.) ........................................................................................................................................................... 178
Figure 7.12 Heat Transfer Coefficient for Different Negative Incidences (Courtesy of Calzada et
al.) 180
Figure 7.13 Stanton Number for Different Negative Incidences (Courtesy of De La Calzada et
al.) 181
Figure 7.14 Stanton Number vs. Reynolds Numbers (Courtesy of De La Calzada et al.) ........... 182
Figure 7.15 Heat Transfer Coefficient vs. Reynolds Number (Courtesy of Calzada et al.)......... 182
Figure 8.1 Pressure Ratio by Normalized Mass Flow (Courtesy of Simoes) ................................... 186
Figure 9.1 Schematics of 3D Concept at IGV/Rotor/Stator Interface................................................. 193
Figure 9.2 Interface Between Rotor/Stator .................................................................................................. 194
Figure 9.3 Difference between Passage and Stages ................................................................................... 194
Figure 9.4 Block Computational Domain for a Rotor with guiding Vanes........................................ 196
Figure 9.5 Axial Rotor/Stator Interaction (Schematics Illustrating the Mixing Plane concepts)
197
Figure 9.6 A Compressor Pressure Distribution on a Surface using a Mixing Plane.................... 197
Figure 9.7 Schematic of an Artificial Interface Between a Rotor and a Stator (left) and the
Virtual Control Volume Formed by Displacing Two Adjacent Domains (right)................................. 198
15

Figure 9.8 Sketch of Casing Treatment of Rotor 67 (Courtesy of 157)................................................. 200


Figure 9.9 Span-Wise Distributions of Aerodynamic Parameters ....................................................... 201
Figure 9.10 Contour of Relative Mach Number and Iso-Surface of Axial Velocity of Modified
Rotor 67 ........................................................................................................................................................................... 201
Figure 9.11 Temperature Contours on the 1st Interface of Modified Rotor 67 - (a) FBEA ; (b)
FBMA ; (c) EA ; (d) MA ; (e) TA (Courtesy of YaLu et al.) ........................................................................ 202
Figure 9.12 Mesh for Transonic Turbine Stage - Upper Image Depicted the Mesh at the Hub
Surface while the Lower Image Represented Mesh used for the Blade Span ..................................... 203
Figure 9.13 Results of the Velocity Contours for a Radial Section at Stator Mid Span using the
Mixing Plane Approach .............................................................................................................................................. 204
Figure 9.14 Schematic View of Pitch-Wise Mixing Model ....................................................................... 205
Figure 9.15 Instantaneous Distributions at 90% Span. ........................................................................... 206
Figure 9.16 Total Pressure Calculated by the Frozen Rotor .................................................................. 207
Figure 9.17 Half Stencil and Full Stencil Reconstruction with: A) Intersection, B) Halo-Cell .. 208
Figure 9.18 Relative Velocities Obtained using HB Techniques ........................................................... 209
Figure 9.19 Phase shifted Periodic Boundary .............................................................................................. 210
Figure 9.20 Phase Shifted Periodic Boundary Conditions ...................................................................... 211
Figure 9.21 Stagnation Pressure Contours under inlet distortion for NASA Rotor 67 ............... 215
Figure 9.22 Computational Mesh for HB and TRS Methods .................................................................. 216
Figure 9.23 Instantaneous Pressure Distribution Within the Compressor Stage Using (NLHB)
216
Figure 9.24 Instantaneous Predictions of Turbulent Viscosity at Mid-Span for HB and TRS
Solutions .......................................................................................................................................................................... 217
Figure 9.25 Non-Dimensional Entropy Generation Using Unsteady (HB) vs Steady (MP) ....... 219
Figure 9.26 Case2: Non-Dimensional Pressure Contours for the Mixing Plane (MP) simulation
(a) , and Harmonic Balance (HB) at Different Time Instances (bcd) ...................................................... 220
Figure 9.27 Mixing Plane vs Harmonic Balance Normalized Entropy Generation Gradients
Obtained with the Adjoint Solution ..................................................................................................................... 222
Figure 9.28 Velocity triangles for the free stream (subscript FS) and the wake (subscript W)
lows. V = absolute velocity, W = relative velocity, U = peripheral velocity .......................................... 226
Figure 9.29 Pattern of entropy evolution (bowing, chopping and transport) of the stator wake
in the rotor channel, as foreseen by CFD ............................................................................................................ 227
Figure 9.30 Simplified Schematic of the Secondary Flows System Downstream of a Rotor .... 228
Figure 9.31 Total pressure loss (Y%), streamwise vorticity (Ωs) and absolute Mach number
(M) downstream of the stator. Experiments at the Fluidmachinery Lab. at Politecnico di Milano
(Italy)................................................................................................................................................................................. 229
Figure 9.32 Rotor inlet low field in the rotating frame of reference. Frame (a) Yloss = total
pressure loss. Frame (b) CPT,R: relative total pressure coefficient. Experiments at the Fluid
machinery Lab. at Politecnico di Milano (Italy). .............................................................................................. 230
Figure 9.33 Schematics of the stator vortical structure transport in the rotor passage ............ 231
Figure 9.34 Time mean flow field downstream of the rotor for a subsonic operating condition
(expansion ratio 1.4, reaction degree at midspan 0.3 and incidence angle close to zero). Frame
(a) relative total pressure coefficient (CPT,R); frame (b) deviation angle (δ). Experiments at the
Fluid machinery Lab. at Politecnico di Milano (Italy). .................................................................................. 231
Figure 9.35 Relative total pressure coefficient on the whole rotor crown. Experiments at the
Fluid machinery Lab. At Politecnico di Milano (Italy). ................................................................................. 232
Figure 9.36 Relative total pressure coefficient (CPT,R), deviation angle (δ) and turbulence (Tu)
for 4 interaction phases. Experiments at the Fluid machinery Lab. at Politecnico di Milano (Italy).
233
Figure 9.37 Standard deviation of the Cptr and δ for the different time frames. Experiments at
16

the Fluid machinery Lab. at Politecnico di Milano (Italy). .......................................................................... 234


Figure 9.38 Rotor loading effects on the stator-rotor interaction. Experiments at the Fluid
machinery Lab. at Politecnico di Milano (Italy). .............................................................................................. 235
Figure 9.39 Vane shock-rotor interaction in axial turbine blades. Red: computation, black:
experiments. Adapted from [Denos et al.] ......................................................................................................... 236
Figure 9.40 Standard deviation for the different instants of the interaction phases. (A) axial
gap: x/bs = 16%; (B) axial gap: x/bs = 35%, nominal; C) axial gap = 50%. Experiments at the
Fluid machinery Lab. at Politecnico di Milano (Italy). .................................................................................. 237
Figure 9.41 Rotor incidence fluctuations in circumferential direction for the different axial
gaps along the blade span ......................................................................................................................................... 238
Figure 9.42 Efficiency trend versus the axial gap. Experiments at the Fluid machinery Lab. at
Politecnico di Milano (Italy). ................................................................................................................................... 238
Figure 10.1 General Description of Computational Planes..................................................................... 241
Figure 10.2 Turbomachine Design Process................................................................................................... 242
Figure 10.3 Optimization Process for Turbomachinery .......................................................................... 243
Figure 10.4 Sketch of a Compressor Stage (left) and Cascade of Geometries at Mid- Span (right)
247
Figure 10.5 Compressor Design Flow Chart ................................................................................................. 248
Figure 10.6 Preliminary Estimation of Number of Stages in Compressor ....................................... 249
Figure 10.7 Optimization Procedures Proposed in [Massardo et al.] ................................................ 250
Figure 10.8 Comparison of Blade Loading Prescribed by Inverse Mode .......................................... 252
Figure 10.9 Turbine Design Process ................................................................................................................ 253
Figure 10.10 Parametric Representation of an Airfoil ............................................................................ 254
Figure 10.11 Sample Mach Number Distribution ....................................................................................... 255
Figure 10.12 Flow path of the turbine ............................................................................................................ 257
Figure 10.13 Schematics of an airfoil showing stream lines along the radial direction ............. 258
Figure 10.14 3D model of an airfoil showing the passage between adjacent airfoils ................. 258
Figure 10.15 Schematic Geometry of theT106D-IZ Turbine Cascade ................................................ 260
Figure 10.16 Shape Optimization History of the Total Pressure Loss Coefficient and
Comparison Between Baseline and Optimized Blade Profile (OptC1)................................................... 261
Figure 10.17 Total Pressure Loss Coefficient Evolution in time Calculated with URANS
Simulation for both the Baseline and the Optimized Configuration ....................................................... 262
Figure 10.18 Mach Number Contours calculated at Three Different Time Instances with the HB
Method, Based on the OptC1 test case, for both the Baseline (a), (b), (c) and the Optimized (d),
(e), (f) Blade Profile ..................................................................................................................................................... 262
Figure 10.19 Tip Speed With Pronounced Negative Camber. From Prince – Courtesy of [Prince]
263
Figure 10.20 Schematic Of Shock Structures (A) Datum, (B) S-Shaped Design. Courtesy of
[John et al.] ...................................................................................................................................................................... 265
Figure 10.21 Datum Geometry and Optimized Shock Control Bumps on The Mid- Section of
Nasa Rotor 67- From Mazaheri et al.. ................................................................................................................. 266
Figure 10.22 The R37 CFD Domain Used – Courtesy of [John et al.] .................................................. 267
Figure 10.23 Radial Profiles Vs Experimental Data – Courtesy of [ John et al.] ............................. 268
Figure 10.24 (a) 3d Separation (Orange) On The R37 Geometry (Flow Right To Left), (b) Rel.
Mach No. Contour At 60% Span – Courtesy of [John et al.]......................................................................... 269
Figure 10.25 Shock Region Flow Features For RR-Fan At Points A) A, B) B, C) C, D) D, E) E, F) F.
Flow Direction – Courtesy of [John et al.] .......................................................................................................... 270
Figure 10.26 Example 2d CST Bump (Solid Line) And The Four Polynomials Used To Construct
It (Dashed Lines) – Courtesy of [John et al.]...................................................................................................... 271
Figure 10.27 a) Example Individual Bump Geometry .............................................................................. 271
17

Figure 10.28 Spanwise Slice Of The Datum And Optimized R37 Geometries At 60% Span –
Courtesy of [John et al.].............................................................................................................................................. 272
Figure 10.29 Optimized R37 Bump (Blue) Added To The Datum Blade Geometry (Grey) ....... 272
Figure 10.30 Datum (Left) And Optimized (Right) Rotor 37 Static Pressure Contours. Flow
Direction Right To Left – Courtesy of [John et al.] .......................................................................................... 273
Figure 10.31 Datum (Left) And R37 Optimized (Right)........................................................................... 274
Figure 10.32 Datum (Left) And Optimized (Right) Rotor 37 Separated Flow Contours (Orange).
Flow Direction Right To Left.................................................................................................................................... 274
Figure 10.33 Lift Plots For The Datum and Optimized............................................................................. 275
Figure 10.34 R37 Optimized Characteristic Vs Datum – Courtesy of [John et al.] ........................ 275
Figure 10.35 Meridional geometry of NASA rotor 37, from AGARD................................................... 279
Figure 10.36 Measurement Stations within NASA Rotor 37, from AGARD ..................................... 280
Figure 10.37 Parameterization of a Compressor Airfoil .......................................................................... 281
Figure 10.38 Multiblock Grid Used in the Simulations ............................................................................. 282
Figure 10.39 Comparison Between Calculated and Experimental Mach Number Contours at
90% Span and m˙ /m˙ choke = 0.98 :Maximum Mach = 1.576, Minimum Mach = 0, and Contour
Interval = 0.031. ............................................................................................................................................................ 282
Figure 10.40 Blade Geometries of the Baseline and Optimized Configurations ............................ 283
Figure 10.41 Mach Number Contours at 95 and 50 % Span of Baseline and Optimized
Geometries ...................................................................................................................................................................... 285
Figure 10.42 Mach Number Contours on The Suction Surface of the Baseline and Optimized
Blades ................................................................................................................................................................................ 286
Figure 10.43 Different Cases of Blade Twist................................................................................................. 288
Figure 10.44 Mesh Generated by TURBOGRID for the Baseline Case ................................................. 290
Figure 10.45 Performance Curves for the Studied Cases ........................................................................ 292
Figure 10.46 Mach Number Distribution for High Mass Flow Point .................................................. 293
Figure 10.47 Total Pressure Over Span for High Mass Flow points at domain outlet ................ 294
Figure 10.48 Entropy distribution for high mass flow points ............................................................... 294
Figure 10.49 Mach number distribution for low mass flow points..................................................... 295
Figure 10.50 Entropy Distribution for Low Mass Flow Points ............................................................. 296
Figure 11.1 Centrifugal impeller with a highly polished surface likely to improve performance
299
Figure 11.2 Cut-Away View of a Turbocharger showing the Centrifugal Compressor ............... 300
Figure 11.3 Jet Engine Cutaway Showing the Centrifugal Compressor among others ............... 301
Figure 11.4 Ninety Degree Inward-Flow Radial Turbine Stage ............................................................ 303
Figure 11.5 Outward Flow Radial Turbine .................................................................................................... 304
Figure 12.1 Different Flow (2D, Q3D, and full 3D) ..................................................................................... 307
Figure 12.2 Full Blade Simulation using Harmonic Balanced Method (Courtesy of CD-adapco)
309
Figure 12.3 Transient Blade Row Extensions Enable Efficient Multi-Stage CFD Simulation
(Courtesy of ANSYS.com) .......................................................................................................................................... 311
Figure 12.4 Typical Meshing of a Turbomachinery Stage ....................................................................... 312
Figure 12.5 Multi-Block Grid for the Space Shuttle Main Engine Fuel Turbine ............................. 312
Figure 12.6 Pressure contour plot, 2nd order spatial discretization scheme................................. 316
Figure 12.7 Analysis provided vibration required for flutter analysis .............................................. 318
Figure 12.8 Examples of Nodal Diameter ...................................................................................................... 319
18
19

Preface
This note is intended for all undergraduate, graduate, and scholars of Turbomachinery. It is not
completed and never claims to be as such. Therefore, all the comments are greatly appreciated. In
assembling that, I was influenced with sources from my textbooks, papers, and materials that I
deemed to be important. At best, it could be used as a reference. I also would like to express my
appreciation to several people who have given thoughts and time to the development of this article.
Special thanks should be forwarded to the authors whose papers seemed relevant to topics, and
consequently, it appears here©. Finally I would like to thank my wife, Sudabeh for her understanding
and the hours she relinquished to me. Their continuous support and encouragement are greatly
appreciated.

Ideen Sadrehaghighi
June 2018
20

1 Introduction
1.1 Preliminaries
Fluid mechanics and thermodynamics are the fundamental sciences used for turbine aerodynamic
design and analysis. Several types of fluid dynamic analysis are useful for this purpose. The concept
through-flow analysis is widely used in axial-flow turbine performance analysis. This involves
solving the governing equations for inviscid flow in the hub-to-shroud plane at stations located
between blade rows. The flow is normally considered to be axisymmetric at these locations, but still
three-dimensional because of the existence of a tangential velocity component. Empirical models are
employed to account for the fluid turning and losses that occur when the flow passes through the
blade rows. By contrast, hub-to-shroud through-flow analysis is not very useful for the performance
analysis of radial-flow turbomachines such as radial-inflow turbines and centrifugal compressors.
The inviscid flow governing equations do not adequately model the flow in the curved passages of
radial turbomachines to be used as a basis for performance analysis. Instead, a simplified “pitch-line”
or “mean-line” one-dimensional flow model is used, which ignores the hub-to-shroud variations.
These also continue to be used for axial-flow turbine performance analysis. Computers are
sufficiently powerful today that there is really no longer a need to simplify the problem that much for
axial-flow turbomachinery. More fundamental internal flow analyses are often useful for the
aerodynamic design of specific components, particularly blade rows. These include 2D flow analyses
in either the blade-to-blade or hub to shroud (Through Flow) direction, and Quasi-3D flow analyses
developed by combining those 2D analyses. Wall boundary layer analysis is often used to supplement
these analyses with an evaluation of viscous effects1.
Viscous CFD solutions are also in use for turbines. These are typically 3D flow analyses, which
consider the effects of viscosity, thermal conductivity and turbulence. In most cases, commercial
viscous CFD codes are used although some in-house codes are in use within the larger companies.
Most design organizations cannot commit the dedicated effort required to develop these highly
sophisticated codes, particularly since viscous CFD technology is changing so rapidly that any code
developed will soon be obsolete unless its development continues as an ongoing activity.
Consequently, viscous CFD is not covered here beyond recognizing it as an essential technology and
pointing out some applications for which it can be effectively used to supplement conventional
aerodynamic analysis techniques.
Prediction of the flow through cascades of blades is fundamental to all aspects of turbomachinery
aerodynamic design and analysis. The flow through the annular cascades of blades in any
turbomachine is really a 3D flow problem. But the simpler two-dimensional blade-to-blade flow
problem offers many advantages. It provides a natural view of cascade fluid dynamics to help
designers develop an understanding of the basic flow processes involved. Indeed, very simple two-
dimensional cascade flow models were used in this educational role long before computational
methods and computers had evolved enough to produce useful design results. Today, blade-to-blade
(B2B) flow analysis is a practical design and analysis tool that provides useful approximations to
many problems of interest. Inviscid blade-to-blade flow analysis addresses the general problem of
two-dimensional flow on a stream surface in an annular. Two-dimensional boundary layer analysis
can be included to provide an approximate evaluation of viscous effects. That approach ignores the
effect of secondary flows that develop due to the migration of low momentum boundary layer fluid
across the stream surfaces. Its accuracy becomes highly questionable when significant flow
separation is present. These limitations require particular care when analyzing the diffusing flow in
compressor cascades. They are less significant for analysis of the accelerating flow in turbine
cascades, but designers still must recognize the approximations and limitations involved. Previously,

1 Ronald H Aungier, e-Books, “The American Society of Mechanical Engineers”, (ASME.org).


21

it have been emphasized the influence of the blade surface velocity distributions on nozzle row and
rotor performance. A graph of the blade surface velocity distributions as a function of distance along
the blade surface is often referred to as the blade-loading diagram. The fundamental role of blade
loading diagrams for the evaluation of blade detailed aerodynamic designs was discussed. Blade-to-
blade flow analysis provides a practical method to calculate these blade-loading diagrams. Indeed,
blade-to-blade flow analysis is an essential part of a modern aerodynamic design system.
A Quasi-3D flow analysis employs 2D flow analyses in the hub-to-shroud and blade-to-blade surfaces
to approximate the 3D flow in a blade passage. The fundamental concept is generally credited to Wu2.
The present analysis achieves exceptional computational speed and reliability largely due to its use
of the linearized blade-to-blade flow analysis. But that also imposes some limitations on the method
that are particularly significant for turbines. Its limitation to subsonic or low transonic Mach number
levels excludes a number of turbine applications. As noted, its accuracy is compromised when it is
applied to the rather thick airfoils often used for turbines. It certainly could be extended for more
general use on turbines by substituting a more general blade-to-blade flow analysis such as the time-
marching method. But that would substantially increase the computation time required and
significantly reduce its reliability. It is very doubtful that this Quasi-3D flow analysis would remain
an attractive design tool if that were done. Indeed, it would lose most of its advantages over
commercially available viscous CFD codes while offering a less general solution. We start with some
explanation of Rotating flow, as well as, derivation of Conservation of Angular Momentum concept
which is fundamental in rotating flows, as well as blade to blade passage.

1.2 Full Engine Simulation Methodology


One of the main challenges in designing an engine is the complexity in terms of geometrical details
(combustion chamber features, turbine cooling holes…) and of interaction effects between the
components which must be modelled with accuracy and acceptable computation time.
Traditionally engine design in
industry has relied on tools like
experimental investigation using
test or flow bench set-ups,
analytical models,
empirical/historical data, 1D/2D
codes and recently, high fidelity 3D
computational fluid dynamics
(CFD) for steady and unsteady flow
physics modelling. Currently most
of the literature work and projects
conducted at an industrial scale
employ a component-by-
component analysis approach,
where each engine component is
studied separately. Such an
approach usually requires
assumptions for inlet and outlet
boundary conditions for each
component and involves a
considerable effort in coupling Figure 1.1 Layout of the KJ66 micro gas turbine
different component analysis tools

2Wu, C. H., "A General Theory of Three Dimensional Flow in Subsonic and Supersonic Turbomachines of Axial,
Radial, and Mixed Flow Types," National Advisory Committee on Aeronautics, NACA TN 2604, 1952.
22

(often with different modelling degrees of accuracy), leading to a process which is potentially error-
prone from the simulation setup point of view and often resulting in significant mismatches between
numerical and experimental data. (See Figure 1.1).
A one-way coupling approach represents one step further in increasing the accuracy of such
simulations. It can be achieved, for example, by extracting outlet profiles of the flow variables from
individual converged component simulations and applying them as inlet boundary condition profiles
to downstream component runs. Nevertheless, the inter-component interaction is still one-way and
the simulation process and results can suffer from similar drawbacks as in the totally uncoupled
workflow. (See Figure 1.2). In the two-way coupling methodology all the components are coupled
and solved simultaneously in one single simulation. This approach greatly simplifies and accelerates
the simulation workflow. Since all the components are considered simultaneously, there is no need
to prescribe boundary conditions between the various elements of the aero-engine. This avoids
running simulations where the states at the interface between the different components have to be
guessed.

Figure 1.2 Three Types of Full Engine Simulation Methodologies


23

2 Preliminary Concepts in Rotating Machinery


2.1 Vortex
One of the major aspects of rotational flow, and in fact the flow in general, is the concept of Vorticity.
In fluid dynamics, a vortex is a region in a fluid in which the flow rotates around an axis line, which
may be straight or curved ( Ting & Klein,
1991)3. The plural of vortex is either
vortices or vortexes. Vortices form in
stirred fluids, and may be observed in
phenomena such as smoke rings,
whirlpools in the wake of boat, or the
winds surrounding a tornado, etc. (see
Figure 2.1). Vortices are a major
component of turbulent flow. The
distribution of velocity, vorticity (the
curl of the flow velocity), as well as the
concept of circulation are used to
characterize vortices. In most vortices,
the fluid flow velocity is greatest next to
its axis and decreases in inverse
proportion to the distance from the axis.
In the absence of external forces,
viscous friction within the fluid tends to
organize the flow into a collection of Figure 2.1 Vortex created by the passage of an aircraft
irrotational vortices, possibly wing, revealed by colored smoke
superimposed to larger-scale flows,
including larger-scale vortices. Once formed, vortices can move, stretch, twist, and interact in
complex ways. A moving vortex carries with it some angular and linear momentum, energy, and
mass4.
2.1.1 Properties of Vortex Flow
2.1.1.1 Vorticity
A key concept in the dynamics of vortices is the vorticity, a vector that describes the local rotary
motion at a point in the fluid, as would be perceived by an observer that moves along with it.
Conceptually, the vorticity could be observed by placing a tiny rough ball at the point in question, free
to move with the fluid, and observing how it rotates about its center. The direction of the vorticity
vector is defined to be the direction of the axis of rotation of this imaginary ball (according to the
right-hand rule) while its length is twice the ball's angular velocity. Mathematically, the vorticity is
defined as the curl (or rotational) of the velocity field of the fluid, usually denoted by ω and expressed
by the vector analysis formula ∇ × u , where u is the local flow velocity. The local rotation measured
by the vorticity ω must not be confused with the angular velocity vector Ω of that portion of the
fluid with respect to the external environment or to any fixed axis. In a vortex, in particular, ω may
be opposite to the mean angular velocity vector of the fluid relative to the vortex's axis.

𝛚= ∇×𝐮=0
Eq. 2.1

3 Ting , L., & Klein, R. (1991). Viscous Vortical Flows (Lecture Notes in Physics). Springer.
4 From Wikipedia, the free encyclopedia.
24

2.1.1.2 Vortex Types


In theory, the speed u of the particles (and, therefore,
the vorticity) in a vortex may vary with the distance r
from the axis in many ways. There are two important
special cases, however:
2.1.1.2.1 Rigid-Body Vortex
If the fluid rotates like a rigid body, that is, if the
angular rotational velocity Ω is uniform, so that u
increases proportionally to the distance r from the
axis. A tiny ball carried by the flow would also rotate
about its center as if it were part of that rigid body (see
Figure 2.2). In such a flow, the vorticity is the same
everywhere, its direction is parallel to the rotation
axis, and its magnitude is equal to twice the uniform Figure 2.2 Rigid-Body Vortex
angular velocity Ω of the fluid around the center of
rotation.

Ω = (0, 0, Ω) , r = (x, y, 0)
Eq. 2.2
u = Ω  r = (-Ω y , Ωx , 0) → ω =   u = (0, 0, 2Ω) = 2Ω
2.1.1.2.2 Irrotational Vortex
If the particle speed u is inversely proportional to the distance r from the axis, then the imaginary
test ball would not rotate over itself; it would maintain the same orientation while moving in a circle
around the vortex axis. In this case the vorticity is zero at any point not on that axis, and the flow is
said to be irrotational.

Ω = (0, 0, r -2 ) , r = (x, y, 0)
Eq. 2.3
u = Ω  r = (- yr -2 ,  xr -2 , 0) → ω =   u = 0

2.1.2 Vortex Geometry


In a stationary vortex, the typical streamline (a line that is everywhere tangent to the flow velocity
vector) is a closed loop surrounding the axis; and each vortex line (a line that is everywhere tangent
to the vorticity vector) is roughly parallel to the axis. A surface that is everywhere tangent to both
flow velocity and vorticity is called a vortex tube. In general, vortex tubes are nested around the axis
of rotation. The axis itself is one of the vortex lines, a limiting case of a vortex tube with zero diameter.
According to Helmholtz's theorems, a vortex line cannot start or end in the fluid – except
momentarily, in non-steady flow, while the vortex is forming or dissipating. In general, vortex lines
(in particular, the axis line) are either closed loops or end at the boundary of the fluid. A whirlpool is
an example of the latter, namely a vortex in a body of water whose axis ends at the free surface. A
vortex tube whose vortex lines are all closed will be a closed torus-like surface. A newly created
vortex will promptly extend and bend so as to eliminate any open-ended vortex lines. For example,
when an airplane engine is started, a vortex usually forms ahead of each propeller, or the turbofan of
each jet engine. One end of the vortex line is attached to the engine, while the other end usually
stretches out and bends until it reaches the ground. When vortices are made visible by smoke or ink
trails, they may seem to have spiral path lines or streamlines. However, this appearance is often an
25

illusion and the fluid particles are moving in


closed paths. The spiral streaks that are taken to
be streamlines are in fact clouds of the marker
fluid that originally spanned several vortex tubes
and were stretched into spiral shapes by the
non-uniform flow velocity distribution. A simple
mathematical description as depicted in Figure
2.3 and can be manufactured as

x = t sin(t) , y = t cos(t) , z = t
Eq. 2.4
2.1.3 Pressure in Vortex
The fluid motion in a vortex creates a dynamic
pressure (in addition to any hydrostatic
pressure) that is lowest in the core region,
closest to the axis, and increases as one moves
away from it, in accordance with Bernoulli's
Principle. One can say that it is the gradient of Figure 2.3 3D Visualization of a Vortex Curve
this pressure that forces the fluid to follow a
curved path around the axis. In a rigid-body vortex flow of a fluid with constant density, the dynamic
pressure is proportional to the square of the distance r from the axis. In a constant gravity field, the
free surface of the liquid, if present, is a concave
paraboloid.
In an irrotational vortex flow with constant fluid
density and cylindrical symmetry, the dynamic
pressure varies as P∞ − K/r2, where P∞ is the limiting
pressure infinitely far from the axis. This formula
provides another constraint for the extent of the core,
since the pressure cannot be negative. The free surface
(if present) dips sharply near the axis line, with depth
inversely proportional to r2. The shape formed by the
free surface is called a hyperboloid. The core of a vortex
in air is sometimes visible because of a plume of water
vapor caused by condensation in the low pressure and
low temperature of the core; the spout of a tornado is
an example. When a vortex line ends at a boundary
surface, the reduced pressure may also draw matter
from that surface into the core. For example, a dust
devil is a column of dust picked up by the core of an air
vortex attached to the ground. A vortex that ends at the
free surface of a body of water (like the whirlpool that
often forms over a bathtub drain) may draw a column
of air down the core (see Figure 2.4). The forward Figure 2.4 A Plughole Vortex
vortex extending from a jet engine of a parked airplane
can suck water and small stones into the core and then into the engine.

2.2 Impeller
An impeller (also written as impellor ) is a rotor used to increase (or decrease in case of turbines) the
pressure and flow of a fluid. It has been used in variety of everyday equipment such as pumps,
26

compressors, medical devices, mixing tanks, water jets and washing machines. More specifically, an
impeller is a rotating component equipped with vanes or blades used in turbomachinery (e. g.
centrifugal pumps). Flow deflection at the impeller vanes allows mechanical power (energy at the
vanes) to be converted into pump power output. Depending on the fluid flow pattern in multistage
pumps and the impellers' arrangement on the pump shaft, impeller design and arrangements are
categorised as: single-stage, multistage, single-entry, double-entry, multiple-entry, in-line (tandem)
or back-to-back arrangement. Axial and radial flow impellers are rotating industrial mixer
components designed for various types of mixing. Both types of impellers are primarily constructed
from stainless steel. Impellers impart flow. They serve the purpose of transferring the energy from
the motor to the substance of a tank as efficiently as possible. Impellers are organized by their flow
patterns.
2.2.1 Types of Impeller
The impeller of a Centrifugal Pump can be of three types as shown in Figure 2.5 as discussed by
• Open Impeller : where the vanes are cast free on both sides.
• Semi-Open Impeller : when the vanes are free on one side and enclosed on the other.
• Enclosed Impeller : The vanes are located between the two discs, all in a single casting 5.

Figure 2.5 Types of Impeller

2.2.2 Flow Characteristics for Impeller


Impellers can be designed to impart various flow characteristics to pump or tank media. Impeller
flow designs can take on three distinct types: Axial, Radial, or Mixed (see Figure 2.7). Because
centrifugal pumps are also classified in this manner, the impeller selection depends upon matching
the pump's flow characteristic to that of the impeller6.

• Axial flow impellers move media parallel to the impeller.


• Radial flow impellers move media at right angles to the impeller itself.
• Mixed flow impellers have characteristics of both axial and radial flow. They may move media
at an angle which is different from right angle radial flow.

5 Presented by: Matt Prosoli, “Centrifugal Pump Overview”, Pumps Plus Inc.
6 See Previous.
27

An impeller is a rotating component of a centrifugal pump,


usually made of iron, steel, bronze, brass, aluminum or
plastic, which transfers energy from the motor that drives
the pump to the fluid being pumped by accelerating the fluid
outwards from the center of rotation. The velocity achieved
by the impeller transfers into pressure when the outward
movement of the fluid is confined by the pump casing.
Impellers are usually short cylinders with an open inlet
(called an eye) to accept incoming fluid, vanes to push the
fluid radially, and a splined, keyed, or threaded bore to accept
a drive-shaft. The impeller made out of cast material in many
cases may be called rotor, also. It is cheaper to cast the radial
impeller right in the support it is fitted on, which is put in
motion by the gearbox from an electric motor, combustion Figure 2.7 A centrifugal pump
engine or by steam driven turbine. The rotor usually names uses an impeller with backward-
both the spindle and the impeller when they are mounted by swept arms
bolts.

Figure 2.6 Flow Direction of Three Different Pumps/Impellers (Courtesy of Global Spec)

2.2.3 Mixing Tanks


Impellers in mixing tanks are used to mix
fluids or slurry in the tank. This can be used
to combine materials in the form of solids,
liquids and gas. Mixing the fluids in a tank is
very important if there are gradients in
conditions such as temperature or
concentration. Figure 2.8 shows two types
of impeller used in mixing tanks, namely:
• Axial flow impeller
• Radial flow impeller
Radial flow impellers impose essentially
shear stress to the fluid, and are used, for
example, to mix immiscible liquids or in
Figure 2.8 Axial Flow Impeller (left) and Radial
general when there is a deformable
Flow Impeller (right)
interface to break. Another application of
28

radial flow impellers are the mixing of very viscous fluids. Axial flow impellers impose essentially
bulk motion, and are used on homogenization processes, in which increased fluid volumetric flow
rate is important. Impellers can be further classified principally into three sub-types7

• Propellers
• Paddles
• Turbines
2.2.4 Axial Impellers
Are best for mixing applications that require stratification or solid suspension. Axial impellers are set
up to create effective top to bottom motion in the tank. This motion is highly effective when placed
over the center of a baffled tank. Some common types of axial flow impellers include: marine
impellers, pitched blade impellers, and hydrofoils. Hydrofoil impellers are also known as high
efficiency impellers. They are a popular choice for applications that require a range from general
blending to storage tanks. This is largely due to the greatest pumping per horsepower, cost
effectiveness, and are ideal for shear sensitive applications.
2.2.5 Radial Impellers
are designed in 4-6 blades. In radial flow impellers, the fluid moves perpendicularly to the impeller.
They produce a radial flow pattern which moves the contents of the mixing tank to the sides of the
vessel. The radial flow impacts the side which causes in either an up or down direction which fills the
top and the bottom of the impeller to be ejected once more. It is also important to note that setting
up baffles helps to minimize vortex and swirling motions in the tank, therefore, enhancing agitation
efficiency. Radial impellers are a great fit for low-level applications inside longer tanks based upon
the production of higher shear due to the angle of attack.
2.2.6 Power Number for Impeller
Power number is a value specific to mixing impellers which describes the impeller's power
consumption. The formula for calculating an impeller's power number is

p
Np = Eq. 2.5
n D 5ρ
3

where Np = power number, P = impeller power in watts, ρ = density of tank liquid in kg/m3, n = shaft
speed in revolutions/second and D=impeller diameter in meters. Because of the difficulty in
obtaining many of these values, power numbers can be considered the summary of various
correlated test results (when dealing with standard-sized mixing tank) rather than a precise
specification. Therefore, manufacturers often specify an impeller's power number as a function of its
power and size8.

2.3 Pumps
A pump is a device that moves fluids (liquids or gases), or sometimes slurries, by mechanical action.
Pumps can be classified into three major groups according to the method they use to move the fluid:
direct lift, displacement, and gravity pumps. Pumps operate by some mechanism (typically
reciprocating or rotary), and consume energy to perform mechanical work by moving the fluid.
Pumps operate via many energy sources, including manual operation, electricity, engines, or wind
power, come in many sizes, from microscopic for use in medical applications to large industrial
pumps. Mechanical pumps serve in a wide range of applications such as pumping water from wells,

7 From Wikipedia, the free encyclopedia.


8 Engineering 360 powered by IEEE global spec.
29

aquarium filtering, pond filtering and aeration, in the car industry for water-cooling and fuel
injection, in the energy industry for pumping oil and natural gas or for operating cooling towers. In
the medical industry, pumps are used for biochemical processes in developing and manufacturing
medicine, and as artificial replacements for body parts, in particular the artificial heart and penile
prosthesis. (see Figure 2.9),
• Single stage pump : When in a casing only one impeller is revolving then it is called single
stage pump.
• Multi stage pump: When in a casing two or more than two impellers are revolving then it is
called double/multi stage pump.
Pumps are used throughout society for a variety of purposes. Early applications includes the use of
the windmill or watermill to pump water. Today, the pump is used for irrigation, water supply,
gasoline supply, air conditioning systems, refrigeration (usually called a compressor), chemical
movement, sewage movement, flood control, marine services, etc. Because of the wide variety of
applications, pumps have a plethora of shapes and sizes: from very large to very small, from handling
gas to handling liquid, from high pressure to low pressure, and from high volume to low volume.

Single Stage Multi Stage

Figure 2.9 Centrifugal Pumps

2.3.1 Types of Pumps


Pump types can be characterized as :
• Positive Displacement Pumps
➢ Rotary Positive Displacement Pumps
➢ Reciprocating Positive Displacement Pumps
➢ Various Positive Displacement Pumps
❖ Gear Pump
❖ Screw Pump
❖ Progressing cavity Pump
❖ Roots-type Pumps
❖ Peristaltic pump
❖ Plunger pumps
❖ Triplex-style plunger pumps
30

❖ Compressed-air-powered double-diaphragm pumps


❖ Rope pumps
• Impulse Pumps
➢ Hydraulic ram pumps
• Velocity pumps
➢ Radial-flow pumps
➢ Axial-flow pumps
➢ Mixed-flow pumps
➢ Jet pump
• Gravity Pumps
• Steam Pumps
• Valve less Pumps

2.3.2 Axial-Flow Pumps vs. Centrifugal Pumps


Axial flow pumps differ from radial flow in that the fluid enters and exits along the same direction
parallel to the rotating shaft. The fluid is not accelerated but instead "lifted" by the action of the
impeller. They may be likened to a propeller spinning in a length of tube. Axial flow pumps operate
at much lower pressures and higher flow rates than radial flow pumps. A centrifugal pump is a roto-
dynamic pump that uses a rotating impeller to increase the pressure and flow rate of a fluid.
Centrifugal pumps are the most common type of pump used to move liquids through a piping system.
The fluid enters the pump impeller along or near to the rotating axis and is accelerated by the
impeller, flowing radially outward or axially into a diffuser or volute chamber, from where it exits
into the downstream piping system. Centrifugal pumps are typically used for large discharge through
smaller heads.

2.4 Some Physics on Rotating Disks Flow


In order to investigate the fluid flow in rotating frames, researchers performed various experiments.
The basic idea is that the (viscous) fluid is confined between two rotating disks9. In general two
boundary layers may be present. The problem is that the equations of motion are so complex, that no
exact solutions are known for this problem even in the stationary regime (one disk fixed the other
rotating). Therefore scientist have to make use of numerical simulations and various experiments to
shed light on the physical mechanisms going on in the rotating fluid10.
2.4.1 Experimental Set-Up
In order to study the flow between
two rotating disks the experimental
set-up shown in Figure 2.10 was
built. The cell consists of a cylinder
of small height h closed by a top disk
and a bottom disk, both of radius
R=140 mm. The upper disk is made
of glass and rotates together with
the cylindrical sidewall which is
made of PVC. The reason why the
cylinder and top disk are made of
PVC and glass is to allow Figure 2.10 Sketch of the experimental set-up
visualization from above and from

9 Miha Meznar, “Fluid Flows In Rotating Frames”, University of Ljubljana, March 2005.
10 Here the focus is not on the recirculation flow but rather on the instability patterns in rotating fluids.
31

side. The bottom disk is made of rectified brass, with a black coating to improve visualization
contrast. To allow the differential rotation the radius of the bottom disk is slightly smaller (a tenth of
millimeter) than the radius of the shrouding cylinder. The thickness h of the cell can be varied
between few mm up to several The cell is filled with a mixture of water, glycerol and small anisotropic
flakes. The latter enable us to visualize the fluid flow. The flakes' orientation with the fluid leads to
variations of the reflected light. For example, if the flakes are mainly horizontal, they reflect light, if
they are vertical they do not reflect it so well. The kinematic viscosity ν = μ/ρ lies between 1x10−6 <
ν < 8 x10−6 m2/s due to different concentration of glycerol11. Each of two disks rotate with its own
angular velocity Ωi, where index i = b, where t stands for bottom and top disk respectively. Angular
velocities of the disks range from 0 to 10 rad/s but the upper disk rotates anticlockwise only, whereas
the bottom one can rotate clock- or anticlockwise. Anticlockwise rotation is taken positive. We call
co-rotation the situation where both disks rotate in the same direction (b and t are of the same sign)
and counter-rotation when the disks rotate in the opposite directions (they have opposite signs). If
one of the disks is left fixed, the other rotating, the regime is called rotor-stator regime. We will define
some dimensionless numbers that describe our cell. The first is radius-to-height ratio defined as Γ=
R/h , where R is radius and h height of the cell. The second number is Reynolds number Rei = Ωih2/ν
, where index i = b, t denotes the bottom and top disk respectively, i is the angular velocity of the disks
and ν the kinematic viscosity. The last number is rotation ratio defined as s = Ωb/Ωt = Reb/Ret .
Rotation ratio is positive (s > 0) in the co-rotation regime and negative (s < 0) in the counter-rotation
regime.
2.4.1.1 Recirculating Flow
Each rotation is associated with a meridian recirculating flow, which can be inward or outward
depending on the rotation ratio. For arbitrary positive and small negative rotation ratio s, the radial
recirculating flow is roughly the same as in the rotor-stator case (s = 0): it consists of an outward
boundary layer close to the faster disk and an inward boundary layer close to the slower disk. At
small negative rotation ratio the centrifugal effect of the slower disk is not strong enough to
counteract the inward flow from the faster disk. But as the rotation ratio s is decreased below −0.2,
the slower disk induces a centrifugal flow too, and the radial recirculating flow appears to come
organized into two-cell recirculating structure. At the interface of these two cells a strong shear layer
takes place. The centrifugal flow induced by the faster disk recirculates towards the center of the
slower disk due to the lateral end wall. This inward recirculation flow meets the outward radial flow
induced by the slower disk, leading to a stagnation circle where the radial component of the velocity
vanishes.
2.4.1.2 Instability Flow Patterns
We now turn to the instability patterns of the flow between two rotating disks close to each other (Γ
= 20.9), in both co- and counter-rotating flows. For s ≥ 0 (rotor-stator or co-rotation) and Reb fixed,
on increasing Ret, propagating circular structures are first observed. These axisymmetric vortices
appear close to the landrail wall, propagate towards the center and disappear before reaching the
center of the cell. Above a secondary threshold of Ret, spiral structures appear at the periphery of the
disks, and circles remain confined between two critical radii (Figure 2.11 (a)). These spirals are
called positive spirals (denoted S+) since they roll up to the center in the direction of the faster disk
(here the top one). Increasing Ret further, positive spirals progressively invade the whole cell. Still
increasing Ret, the flow becomes more and more disordered (denoted D, Figure 2.11 (c)). It can be
shown that co-rotation shifts upwards the instability thresholds for circles and positive spirals12.
However, threshold line for circles is parallel to the solid body rotation (b = t) indicating that the

11 G. Gauthier, P. Gondret, F. Moisy and M. Rabaud, “Instabilities in the flow between co- and counter-rotating
disks”, J. Fluid Mech, volume 473, pp. 1-21, 2002.
12 Miha Meznar, “Fluid Flows In Rotating Frames”, University of Ljubljana, March 2005.
32

Figure 2.11 For s ≥ 0 co-rotation at different speed

angular velocity difference Ω = Ωt − Ωb is the only control parameter of this instability and no influence
of the global rotation occurs. By contrast, the borderline for the positive spirals has a larger slope
than the solid body rotation line; in this case the relative angular velocity Ω is not the only control
parameter and an extra velocity of the upper disk is needed for the spirals to arise. The global rotation
in this case has a stabilizing effect.
For s < 0 (counter-rotating case) the onset of the instability patterns depends on the Reynolds
numbers of both disks. For low bottom Reynolds number, −11 < Reb < 0, on increasing the Reynolds
number of the upper disk, the appearance of the instability patterns is the same as in the rotor-stator
or co-rotation case: axisymmetric propagating vortices, positive spirals and disorder. But, for −18 <
Reb < −11, spirals of a new kind appear on increasing Ret. These spirals are said to be negative (and
denoted S−) since they now roll up to the center in the direction of the slower counter-rotating disk
(Figure 2.12 (a)). Unlike circles and positive spirals, negative spirals extend from the periphery to
the center, they invade the whole cell. Also, the onset time for negative spirals is much longer than
for positive ones or circles; when the onset is carefully approached from below, the growth time of
negative spirals can exceed 15 minutes which strongly contrasts circles and positive spirals which
appear almost instantaneously. Increasing Ret further, positive spirals appear as well at the periphery
of the disk, as can be seen in Figure 2.12 (b). Here negative and positive spirals seem to coexist
without strong interaction, which indicates the difference in their origin. The circles and positive

Figure 2.12 For s < 0 counter-rotating at different speed


33

spirals have their origin in the boundary layer instability whereas negative ones, on the other hand,
originate from shear layer instability.
Still increasing Ret, negative spirals disappear and positive spirals alone remain (Figure 2.12(c)).
Increasing Ret yet further, circles appear as in the co-rotation case. Still increasing Ret, the structures
become disorganized and the flow becomes turbulent. For Reb < −18 the negative spirals described
above become wavy, the flow is more and more disorganized and continuously becomes turbulent
without a well-defined threshold. Depending on the Reynolds number, the disorder can be generated
first at the periphery or in the center and then invades the entire cell. Up to now our instability
patterns were limited to radius-to-height ratio Γ = 20.9. Does anything changes if one changes it?
Researchers enlarged the gap h between the disks (Γ diminishes) and observed a new pattern that
consisted of a sharp-cornered polygon of m sides, surrounded by a set of 2m outer spiral arms. These
polygons arise only for small Γs (less than approx. Γ = 10). For higher values the vertical confinement
leads to a saturated pattern where inner arms, connecting the corners of the polygon to the center of
polygons, turn into negative spirals. Another interesting property of the patterns is that they are not
fixed but rather rotate as a whole. Therefore we define the azimuthal phase velocity ωφ in the
laboratory frame. It corresponds to the angular velocity of the global rotation of the spiral pattern.
For the S+ spirals ωφ is always positive (anticlockwise), i.e. the positive spirals rotate in the direction
of the faster (top) disk, regardless of motion of the bottom one. S− spirals, on the other hand change
sign of ωφ. It means that for small Ret the pattern rotates in the direction of the slower (bottom) disk
while at higher Ret it moves with the top (faster) disk. Here only compare the directions of the disks
and phase velocity. The size of phase velocity is only a fraction of the disk velocities. We see that the
co-rotation flow (Reb > 0, right-hand part of the diagram) is qualitatively the same as the rotor-stator
flow (vertical line Reb = 0); the thresholds of instabilities (circles C and positive spirals S+) are found
to increase just with the bottom Reynolds number. By contrast, the counter-rotating case (Reb < 0,
left-hand part) is much more rich.

2.5 Discussion on Effects Swept and Dihedral Blades


We have encountered difficulty with the terminology that describes different blade stacking lines
because sweep and dihedral have distinct aerodynamic effects. It is convenient to reserve the
meaning of dihedral to be lean at right angles to the sweep so the translation producing it is normal
to the stagger line. As blades, described as having "dihedral," actually has aft sweep as well as
dihedral. It would seem
more logical, as well as
more desirable, to
describe the DHP blades
as having "tangential
lean," which includes
both true dihedral and
sweep, (see Figure
2.13).
Over the forward part of
the blade, the forward
sweep induces flow near
the suction surface
toward the end-wall and
induces flow near the
pressure surface away
from the end-wall. These
changes to the span-wise Figure 2.13 Geometry of Swept and Dihedral Blades
flow oppose the classical
34

Figure 2.14 Experimental Pressure Iso-Surfaces ; Left - Without Sweep ;Right - With Forward Sweep
(Courtesy of RÁBAI and VAD [93])
secondary flow created by turning the incoming end-wall boundary layer within the blade passage,
and therefore reduce the cross-passage flow of end-wall fluid toward the suction surface. Aft sweep
will tend to add to the classical secondary flow and there is then more loss and blockage buildup on
the suction surface.
Forward sweep leads to enlargement of the meridional stream-tube near the end-wall ahead of the
leading edge followed by a contraction and then again an opening further downstream. The sequence
leads to a corresponding acceleration,
deceleration, and acceleration of the end-wall
flow relative to that for un-swept blades; the
consequence is usually lower end-wall loss (see
Figure 2.14)13. Aft sweep has the opposite effects
and leads to an increase in end wall loss. Dihedral,
on the other hand, causes the bound vortex and its
image to induce velocities more or less parallel to
the primary flow. If the dihedral is positive, so that
the suction surface makes an copious angle with
the end-wall, the induced velocities tend to reduce
the velocity peak on the suction surface, which
reduces the diffusion near the suction surface
end-wall corner.

2.6 Linear Cascade


Flow within the blade passages of an axial
turbomachine is three-dimensional. A two-
dimensional approximation is obtained by
making a cylindrical section of the blade row at
the mean radius and by unrolling it into a plane
(see Figure 2.15). A linear cascade or blade row
is built by setting up prismatic blades with the Figure 2.15 Linear Turbine Cascade

13Gergely Rábai And János Vad, “Validation Of A Computational Fluid Dynamics Method To Be Applied To Linear
Cascades Of Twisted-Swept Blades”, Periodica Polytechnica Ser. Mech. Eng. Vol. 49, No. 2, Pp. 163–180 (2005).
35

profiles obtained. In principle, an infinite number of blades are required in order to keep the
periodicity of the blades in the original machine. The two-dimensional flow is representative for the
real three-dimensional flow. The expressions for axial momentum are the same in both
configurations. Moment of momentum with the three-dimensional flow corresponds to tangential
momentum with the cascade (tangential = circumferential)14.
Constant tangential momentum along the blade height with the linear cascade is then equivalent to
constant angular momentum (vur = constant) along the radius in the real machine. This means then
that the three-dimensional flow is a so-called free vortex flow. The term means swirling flow with
constant angular momentum.

14Erik Dick, “Fundamentals of Turbomachines”, Springer Dordrecht Heidelberg New York London. ISBN 978-
94-017-9626-2.
36

3 Conservation of Angular Momentum & Rotating Reference Frame


3.1 Flow in Rotating Reference Frame
Consider a coordinate system which is rotating steadily with angular velocity ω (bold face represents
the vector quantity in picture) relative to a stationary (inertial) reference frame, as illustrated in
Figure 3.1. The origin of the rotating system is located by a position vector r015. The fluid velocities
can be transformed from the stationary frame to the rotating frame using the following relation:

𝐮𝐫 = 𝐮 − 𝛚
⏟×𝐫 where 𝛚 = 𝛚 𝐚̂
whirl velocity
Eq. 3.1
In the above, ur is the relative velocity (the velocity viewed from the rotating frame), u is the
absolute velocity (the velocity viewed from the stationary frame), ω x r is the whirl (or moving)
velocity (the velocity due to the moving frame), and â is unit directional vector depending or rotation
direction. When the equations of motion are solved in the rotating reference frame, the acceleration
of the fluid is augmented by additional terms that appear in the momentum equations (Batchelor,
2000)16. Moreover, the equations can be formulated in two different ways:
• Expressing the momentum equations using the relative velocities as dependent variables
(known as the relative velocity formulation).
• Expressing the momentum equations using the absolute velocities as dependent variables in
the momentum equations (known as the absolute velocity formulation).
The exact forms of the governing equations for these two formulations will be provided in the
sections below. It can be noted here that pressure-based solvers provide the option to use either of
these two formulations, whereas the density-based solvers always use the absolute velocity
formulation.

Figure 3.1 Rotating Frame of Reference

15FLUENT 6.3 User's Guide.


16Batchelor, G. (2000). An Introduction to Fluid Dynamics (Cambridge Mathematical Library). Cambridge:
Cambridge University Press. doi:10.1017/CBO9780511800955
37

3.1.1 Centrifugal & Coriolis Forces


From elementary dynamics, the centrifugal force is an inertial force (also called a 'fictitious' or
'pseudo' force) directed away from the axis of rotation that appears to act on all objects when viewed
in a rotating reference frame (i.e., ω = curl v). The Coriolis force is an inertial force (also called a
fictitious force) that acts on objects that are in motion relative to a rotating reference frame (Apsley,
2017)17. In reference frame with clockwise rotation, the force acts to the left of the motion of the
object. In one with anticlockwise rotation, the force acts to the right (see Figure 3.2).

ωω

Figure 3.2 Centrifugal and Coriolis Force

3.1.2 Relative Velocity Formulation


For the relative velocity formulation, the governing equations of fluid flow for a steadily rotating
frame can be written as follows

∂ρ
Mass + ∇. (ρ𝐮r ) = 0
∂t

(ρ𝐮r ) + ∇. (ρ𝐮r 𝐮r ) + ρ (2𝛚
⏟ × 𝐮r + ⏟ 𝛚 × 𝛚 × 𝐫) = −∇p + ∇𝛕rij + 𝐅
∂t Centrifugal
Coriolis

Energy (ρEr ) + ∇. (ρ𝐮r Hr ) = ∇. (k∇T + 𝛕rij ur ) + 𝐒h
∂t
p 1 2 p
Er = h − + [𝐮r − (𝛚 × 𝐫)2 ] , Hr = Er +
ρ 2 ρ
Eq. 3.2
Here, the subscript r is for relative quantities, the momentum equation contains two additional
acceleration terms, the Coriolis acceleration (2ω x ur), and the Centrifugal acceleration (ω x ω x r).
In addition, the viscous stress τijr is defined as before except that relative velocity derivatives are
used. The energy equation is written in terms of the relative internal energy (Er) and the relative total

17 David Apsley, “Fluid-Flow Equations”, spring 2017.


38

enthalpy (Hr), also known as the rothalpy18.


3.1.3 Absolute Velocity Formulation
For the absolute velocity formulation, the governing equations of fluid flow for a steadily rotating
frame can be written as follows:

∂ρ
Mass + ∇. (ρur ) = 0
∂t

Momentum (ρu) + ∇. (ρur u) + ρ (ω
⏟ × u) = −∇p + ∇τij + F
∂t I

Energy (ρE) + ∇. (ρur H + p(ω × r) = ∇. (k∇T + τij . u) + Sh
∂t
Eq. 3.3
In this formulation, the Coriolis and Centripetal accelerations can be collapsed into a single term (I).
Be advised that from now on we will be dealing with linear momentum if noted otherwise19.
3.1.4 Early Formulation and Consideration
There are more simplified version of equations depending to the type of analysis. One particular
version is axisymmetric which usually governs the aerospace cruse condition. For example a time-
marching finite volume numerical procedure is presented for 3D Euler analysis of turbomachinery
flows by [Soulis]20. Another pioneering and early studies in the subject prediction of turbomachinery
performance, as indicated by the flow in the through-flow, or axisymmetric plane, is done by [Siebert
and Yocum]21 when only steady-state and incompressible flows are considered.

3.2 Flows with Rotating Reference Frames


By default, the equations of fluid flow and heat transfer are solves in a stationary (or inertial)
reference frame. However, there are many problems where it is advantageous to solve the equations
in a moving (or non-inertial) reference frame. Such problems typically involve moving parts (such
as rotating blades, impellers, and similar types of moving surfaces), and it is the flow around these
moving parts that is of interest. In most cases, the moving parts render the problem unsteady
when viewed from the stationary frame. With a moving reference frame, however, the flow
around the moving part can (with certain restrictions) be modeled as a steady-state problem with
respect to the moving frame. The moving reference frame modeling capability allows you to model
problems involving moving parts by allowing you to activate moving reference frames in selected cell
zones. When a moving reference frame is activated, the equations of motion are modified to
incorporate the additional acceleration terms which occur due to the transformation from the
stationary to the moving reference frame. By solving these equations in a steady-state manner, the
flow around the moving parts can be modeled. For simple problems, it may be possible to refer the
entire computational domain to a single moving reference frame. This is known as the Single
Reference Frame (SRF) approach. The use of the SRF approach is possible, provided the geometry
meets certain requirements22. For more complex geometries, it may not be possible to use a single

18 FLUENT 6.3 User's Guide.


19 Same as previous.
20 Johannes Vassiliou Soulis, “An Euler Solver For Three-Dimensional Turbomachinery Flows”, International

Journal For Numerical Methods In Fluids, Vol. 20,L-30, 1995.


21 B. W. Siebert, A.M. Yocum, “An Incompressible Axisymmetric Through-Flow Calculation Procedure For Design

And Off-Design Analyses Of Turbomachinery”. Technical Report No. Tr 93-05, 1993.


22 FLUENT 6.3 User's Guide.
39

reference frame. In such cases, you must break up the problem into multiple cells zones, with well-
defined interfaces between the zones (MRF). The manner in which the interfaces are treated leads
to three approximate, steady-state modeling methods for this class of problem:
• Single Reference Frame (SRF)
• Multiple Reference Frame (MRF)
• Mixing Plane Method (MPM)
If unsteady interaction between the stationary and moving parts is important, you can employ the
Sliding Mesh approach, or more simplified Non-Uniform Harmonic Balanced Method, to capture
the transient behavior of the flow.
3.2.1 Single Rotating Reference Frame (SRF) Modeling
Many problems permit the entire computational domain to be referred to as a single rotating
reference frame (SRF modeling). In such cases, the equations for a Rotating Reference Frame are
solved in all fluid cell zones. Steady-state solutions are possible in SRF models provided suitable
boundary conditions are prescribed. In particular, wall boundaries must adhere to the following
requirements:
• Any walls which are moving with the reference frame can assume any shape. An example
would be the blade surfaces associated with a pump impeller. The no slip condition is defined
in the relative frame such that the relative velocity is zero on the moving walls.
• Walls can be defined which are non-moving with respect to the stationary coordinate system,
but these walls must be surfaces of revolution about the axis of rotation. Here the so slip
condition is defined such that the absolute velocity is zero on the walls. An example of this
type of boundary would be a cylindrical wind tunnel wall which surrounds a rotating
propeller.

Rotationally periodic boundaries may also be used, but the surface must be periodic about the axis
of rotation. As an example, it is very common to model through a blade row on a turbomachine by
assuming the flow to be
rotationally periodic and using a
periodic domain about a single
blade. This permits good
resolution of the flow around the
blade without the expense of
model all blades in the blade row
(see Figure 3.3). Flow boundary
conditions (inlets and outlets) can
be in most cases prescribed in
either the stationary or rotating
frames. For example, for a velocity
inlet, one can specify either the
relative velocity or absolute
velocity, depending on which is
more convenient. In some cases Figure 3.3 Single Blade Model with Rotationally Periodic
(e.g. pressure inlets) there are Boundaries
restrictions based upon the
velocity formulation which has been chosen. For additional information, reader should refer to
FLUENT 6.3® user manual.
40

3.2.1.1 Case Study - Aerodynamics and Structural Analysis of (HAWT) Wind Turbine Blade
Wind blade design is determined using blade element momentum (Thresher & Dodge, 1998)23. The
blade plays a pivotal role, because it is the most important part of the energy absorption system.
Practical horizontal axis wind turbine (HAWT) designs use airfoils to transform the kinetic energy in
the wind into useful energy and it has to be designed carefully to enable to absorb energy with its
greatest efficiency. There are many factors for selecting a profile. One significant factor is the chord
length and twist angle which depend on various values throughout the blade. In this work, the airfoil
sections used in horizontal axis wind turbine (HAWT) are S818; S825 and S826 airfoils used in NREL
phase 2 and phase 3 wind turbines. They have several advantages in meeting the intrinsic
requirements for wind turbines in terms of design point, off-design capabilities and structural
properties. The lift and drag goes beyond the traditional aim of capacity maximization, contributing
also for organization’s profitability and value. Indeed, lean management and continuous
improvement approaches suggest capacity optimization instead of maximization. The study of
capacity optimization and costing models is an important research topic that deserves contributions
from both the practical and theoretical perspectives. This paper presents and discusses a
mathematical model for capacity management based on different costing models (ABC and TDABC).
A generic model has been developed and it was used to analyze idle capacity and to design strategies
towards the maximization of organization’s value. The trade-off capacity maximization vs
operational efficiency is highlighted and it is shown that capacity optimization might hide operational
inefficiency.
3.2.1.1.1 Introduction
The power efficiency of wind energy systems has a high impact in the economic analysis of this kind
of renewable energies. The efficiency in these systems depends on many subsystems: blades,
gearbox, electric generator and control. Some factors involved in blade efficiency are the wind
features, e.g. Its probabilistic distribution, the mechanical interaction of blade with the electric
generator, and the strategies dealing with pitch and rotational speed control 24. It is a complex
problem involving many factors, relations and constraints. The design of optimal blades involves
aerodynamic, structural and control problems. However, the design cycle can be practically
approached as an iterative and stepped method. For aerodynamic optimization the blade can be
modelled as a series of sections along the pitch axis. Each section has an airfoil shape, chord length
and attach angle which is the result of a collective pitch angle and a local twist one. This last is a
property of the blade while the pitch angle depends on the control strategy of the whole energy
system.
The computation of the wind flow around rotating blades is a very complex problem. For a precise
knowledge of the wind flow and the induced forces in the turbine surfaces it is necessary to solve the
three-dimensional Navier-Stokes equations in a rotating frame, but the computational cost to obtain
such precise solution prohibits their use in the design and analysis environments (Thresher & Dodge,
1998). The blade element momentum theory (BEM) is basically a one-dimensional simplified theory
that is used routinely by wind power industry because it provides reasonably accurate prediction of
performance25. The BEM theory has shown to give good accuracy with respect to time cost, and at
moderate wind speeds, it has sufficed for blade geometry optimization 26-27-28. The BEM theory is the

23 Thresher, R. W., & Dodge, D. M. (1998). Trends in the evolution of wind turbine generator configurations and
systems. WIND ENERGY.
24 K.Y. Maalawi, M.A. Badr, A practical approach for selecting optimum wind rotors, Renewable Energy, (2003).
25 J.L. Tangler, The Nebulous art of using wind-tunnel airfoil data for predicting rotor Performance, 5 (2002).
26 M.M. Duquette, K.D. Visser, Numerical implications of solidity and blade number on rotor performance of

horizontal-axis wind turbines, J. Sol. Energy Eng.-Trans, ASME. 125 (2003) 425–432, 2003.
27 P. Fuglsang, H.A. Madsen, Optimization method for wind turbine rotors, J. Wind E, Ind. Aerodynamics. (1999).
28 M. Jureczko, M. Pawlak, A. Mezyk, Optimisation of wind turbine blades, J. Mater, Proc. Technol. 167 (2005).
41

composition of two different approaches to study the forces in a wind turbine. The first is the
momentum theory that studies the global changes in wind momentary, axial and tangential, in an
ideal turbine. Changes in axial and rotational moment between upwind and downwind induce thrust
and torque respectively in the rotor. The wind flow is split in many differential non-interacting
annular stream tubes. The second theory, the blade element, studies the aerodynamic forces acting
in a local airfoil. As in aeronautics wing theory, the forces are lift, which is perpendicular to the wind
direction, and drag that is in the same direction. Drag is mainly generated by friction between the
viscous fluid and the airfoil surface. It is a dissipative force that generates power loss and lack in
momentum changes. The applications of thick airfoils are extended to the assessment of wind turbine
performance. It is well established that the power generated by a Horizontal-Axis Wind Turbine
(HAWT) is a function of the number of blades B, the tip speed ratio λ (blade tip speed/wind free
stream velocity) and the lift to drag ratio (CL/CD) of the airfoil sections of the blade. The airfoil
sections used in HAWT are generally thick airfoils such as the S818, S825 and S826 of airfoils. These
airfoils vary in (CL/CD) for a given B and λ, and therefore the power generated by HAWT for different
blade airfoil sections will vary. Another goal of this study is to evaluate the effect of different airfoil
sections on HAWT performance using the Blade Element Momentum (BEM) theory.
3.2.1.1.2 Airfoils and General Concepts of Aerodynamics
A number of terms are used to characterize an airfoil. The mean camber line is the locus of points
halfway between the upper and lower surfaces of the airfoil. The most forward and rearward points
of the mean camber line are
on the leading edge and
trailing edges, respectively.
The straight line connecting
the leading and trailing
edges is the chord line of the
airfoil, and the distance from
the leading to the trailing
edge measured along the
chord line is designated as
the chord of the airfoil. The Figure 3.4 Airfoil Cross-Sections used in the Design of the Wind
thickness is the distance Turbine Blades
between the upper and
lower surfaces, also
measured perpendicular to the chord line. Finally, the angle of attack α is defined as the angle
between the relative wind and the chord line. (See Figure 3.4)
3.2.1.1.3 Lift, Drag and Non-Dimensional Parameters
Theory and research have shown that many flow problems can be characterized by non-dimensional
parameters. The most important non-dimensional parameter for defining the characteristics of fluid
flow conditions is the Reynolds number. Force and moment coefficients, which are a function of
Reynolds number, can be defined for two or three-dimensional objects. Force and moment
coefficients for flow around two-dimensional objects are usually designated with a lower case
subscript lift and drag coefficients that are measured for flow around two- or three-dimensional
object are usually designated with an upper case subscript. Rotor design usually uses two
dimensional coefficients, determined for a range of angles of attack and Reynolds numbers, in wind
tunnel tests. The two-dimensional lift and drag coefficients is defined as:
42

L/l D/l
CL = , CD =
(1/2)ρU 2 c (1/2)ρU2 c
Eq. 3.4
where ρ is the density of air, U is the velocity of undisturbed airflow, c is the airfoil chord length and
l is the airfoil span. Other dimensionless coefficients that are important for the analysis and design of
wind turbines include the power and thrust coefficients and the tip speed ratio, the pressure
coefficient, which is used to analyze airfoil flow:

p − p∞
Cp =
(1/2)ρU2
Eq. 3.5
Under similar ideal conditions, symmetric airfoils of finite thickness have similar theoretical lift
coefficients. This would mean that lift coefficients would increase with increasing angles of attack
and continue to increase until the angle of attack reaches 90 degrees. The behavior of real symmetric

(a) Lift coefficients (b) Drag coefficients for the


variations vs angle of attack S818 airfoil at Re=1.106

(c) Lift coefficients variations


(d) Drag coefficients for the
vs angle of attack
S825 airfoil at Re=1.106

Figure 3.5 Typical Lift and Drag Coefficients

airfoils does indeed approximate this theoretical behavior at low angles of attack. For example,
typical lift and drag coefficients for a S818, S825, and S826airfoils,the profiles of which are shown in
Figure 3.5 (a-d) as a function of angle of attack and Reynolds number.
43

Note that, in spite of the very good correlation at low angles of attack, there are significant differences
between actual airfoil operation and the theoretical performance at higher angles of attack. The
differences are due primarily to the assumption, in the theoretical estimate of the lift coefficient, that
air has no viscosity. Surface friction due to viscosity slows the airflow next to the airfoil surface,
resulting in a separation of the flow from the surface at higher angles of attack and a rapid decrease
in lift. This condition is referred to as stall. Airfoils for horizontal axis wind turbines (HAWTs) often
are designed to be used at low angles of attack, where lift coefficients are fairly high and drag
coefficients are fairly low. The lift coefficient of this symmetric airfoil is about zero at an angle of
attack of zero and increases to over 1.0 before decreasing at higher angles of attack. The drag
coefficient is usually much lower than the lift coefficient at low angles of attack. It increases at higher
angles of attack. The lift coefficient at low angles of attack can be increased and drag can often be
decreased by using a cambered airfoil.
3.2.1.1.4 Blade Element Momentum (BEM) Theory
The blade element momentum (BEM) theory is a compilation of both momentum theory and blade
element theory (M. Jureczko, M. Pawlak, A. Mężyk, 2005)29 and (Ernesto Benini, 2003)30. Momentum
theory, which is useful in predicted ideal efficiency and flow velocity, is the determination of forces
acting on the rotor to produce the motion of the fluid. Theory determines the forces on the blade as
a result of the motion of the fluid in terms of the blade geometry. By combining the two theories,
BEM theory, also known as strip theory, relates rotor performance to rotor geometry.
3.2.1.1.5 Aerodynamic Load
Aerodynamic load is generated by lift and drag of the blades airfoil section, which is dependent on
wind velocity, blade velocity, angle of attack and yaw31. The angle of attack is dependent on blade
twist and pitch. The aerodynamic lift and drag produced are resolved into useful thrust in the
direction of rotation absorbed by the generator and reaction forces. It can be seen that the reaction
forces are substantial acting in the flat wise bending plane, and must be tolerated by the blade with
limited deformation. For calculation of the blade aerodynamic forces the widely publicized blade
element momentum (BEM) theory is applied. Working along the blade radius taking small elements
δr, the sum of the aerodynamic forces can be calculated to give the overall blade reaction and thrust
loads.
3.2.1.1.6 Blade Geometry
The main objective in the design of wind turbines is to find a rotor that meets the basic conditions
requested. The most important condition is to get a rotor to deliver output power required at a
particular speed. For this, the first assumption of the aerodynamic rotor is its diameter, which can be
roughly estimated power. In addition, it is necessary to take into account the importance of the
geometry of the rotor, taking into consideration the most important, the aerodynamic performance,
strength and stiffness conditions, and costs. However, power generation through wind turbines also
play a decisive role in the design of the aerodynamics of the rotor, which is influenced by other
parameters such as power generator and control system. The results of the optimal distribution of
the cord and twist for a blade of 43.2 m in diameter and having various profiles.

29 M. Jureczko, M. Pawlak, A. Mężyk, Optimisation of wind turbine blades, Journal of Materials Processing
Technology, Volume 167, Issues 2–3, 2005, Pages 463-471, ISSN 0924-0136,
https://doi.org/10.1016/j.jmatprotec.2005.06.055
30 Ernesto Benini, Significance of blade element theory in performance prediction of marine propellers, Ocean

Engineering, Volume 31, Issues 8–9, 2004, Pages 957-974, ISSN 0029-8018,
https://doi.org/10.1016/j.oceaneng.2003.12.001
31 M.O.L. Hansen, Aerodynamics of Wind Turbines: Rotors, Loads and Structure, James & James Science

Publishers, London, UK, pp. 48-59, 2000.


44

3.2.1.1.7 Mathematic Model


The governing equations are the continuity and Navier-Stokes equations. These equations are
written in a frame of reference rotating with the blade. This has the advantage of making our
simulation not require a moving mesh to account for the rotation of the blade. The equations cab be
obtained from (Mouhsine et al., 2017)32.
3.2.1.1.8 Turbulence Model
The SST 𝑘−𝜔 turbulence model33-34 is a two equation eddy-viscosity model which has become very
popular. The shear stress transport (SST) formulation combines the better of two worlds. The use of
a 𝑘−𝜔 formulation in the inner parts of the boundary layer makes the model directly usable all the
way down to the wall through the viscous sub-layer; hence the SST 𝑘 − 𝜔 model can be used as a Low-
Re turbulence model without any extra damping functions. The SST formulation also switches to a 𝑘
−𝜀 behavior in the free-stream and thereby avoids the common 𝑘 − 𝜔 problem that the model is too
sensitive to the inlet free-stream turbulence properties. Authors who use the SST 𝑘−𝜔 model often
merit it for its good behavior in adverse pressure gradients and separating flow. The SST 𝑘−𝜔 model
does produce a bit too large turbulence levels in regions with large normal strain, like stagnation

Figure 3.6 (Top) Airfoils Superposed on the Wind Turbine Blade and (Bottom) Top View of a Subset of
the Airfoil Cross-Sections illustrating Blade Twisting

32 Saaa E. Mouhsine, Karim Oukassou, Mohammed Marouan Ichenial, Bousselham Kharbouch, Abderrahamane
Hajraoui, “Aerodynamics and Structural Analysis of Wind Turbine Blade”, 11th International Conference
Interdisciplinarity in Engineering, INTER-ENG 2017, 5-6 October 2017, Romania.
33 D. C. Wilcox, Formulation of the k–ω Turbulence Model Revisited, 11, AIAA Journal, 823–2838, 2008.
34 F. R. Menter, “Zonal Two Equation k-ω Turbulence Models for Aerodynamic Flows”, AIAA Paper, Vol. 93, 1993.
45

regions and regions with strong acceleration. Details of the SST 𝑘−𝜔 turbulence model is available in
[Mouhsine et al.]35.
3.2.1.1.9 Mesh Generation
In order to create the computational domain and generate mesh, the commercially available software
“ANSYS Meshing tool” is used to build a wind tunnel model and generate an unstructured mesh
around the blade in the computational domain. As shown in Figure 3.6, a 3D straight untampered
blade is placed inside a computational domain (mimicking a wind tunnel) with inflow and outflow
boundaries. The wall boundary condition is applied to the right and lift surface of the computational
domain. The back and front surfaces of the computational box are set as symmetry boundary
condition due to the free motion of air on these surfaces. As shown, one important part of the mesh
shape is that it must be smooth and dense enough to be suitable for any arbitrary airfoil shape and
3D Horizontal-Axis Wind Turbine. The wind tunnel geometry is always the same, but the airfoil/blade
in the center of the tunnel changes from one generation to the next. This poses a challenge for mesh
generation in 3D. Faces are meshed using quadrilateral cells, and we require that the number of
nodes on opposite faces be identical. To ensure this distribution, we define a set number of nodes
(and not relative node spacing) along each edge. Otherwise, thicker or more cambered airfoil edges
would have more nodes than thinner ones if a relative distribution was used. A refined boundary
layer is carefully constructed around the airfoil to capture the boundary layer behavior.

Figure 3.7 Force Analysis for S818 Airfoil Section

35Saaa E. Mouhsine, Karim Oukassou, Mohammed Marouan Ichenial, Bousselham Kharbouch, Abderrahamane
Hajraoui, “Aerodynamics and Structural Analysis of Wind Turbine Blade”, 11th International Conference
Interdisciplinarity in Engineering, INTER-ENG 2017, 5-6 October 2017, Romania.
46

3.2.1.1.10 Result and Discussion


The Horizontal-Axis Wind Turbine (HAWT) is 43.2 meters long and starts with a cylindrical shape at
the root and then transitions to the airfoils S818, S825 and S826 for the root, body and tip,
respectively. This wind blade also has pitch to vary as a function of radius, giving it a twist and the
pitch angle at the blade tip is 4 degrees. Accordingly, the stress limit of the blade is determined by
the strength of the E-glass used in the skin of the blade. The turbulent wind flows towards the
negative z-direction at 12 m/s which is a typical rated wind speed for a turbine at this size. This
incoming flow is assumed to make the blade rotate at an angular velocity of-2.22 rad/s about the z-
axis. The tip speed ratio is therefore equal to 8 which is a reasonable value for a large wind turbine.
The blade root is offset from the axis of rotation by 1 meter. The process of CFD simulation begins
with the creation of a three dimensional domain and its proper discretization. We define the velocity

(a) velocity distribution

(b) Pressure contours at several blade cross-sections at


viewed from the back of the blade.

(c) Full Blade Deformation for Cut-Out Wind Speed

(d) Stress for Cut-Out Wind Speed

Figure 3.8 Static Pressure, Velocity Magnitudes, Deformation and Stress Distribution
47

at the inlet of 12 m/s with turbulent intensity of 5% and turbulent viscosity ratio of 10 and the
Pressure of 1 atm in order to validate the present simulation. As mentioned in the beginning of this
work, the aerodynamic performance of wind turbines are primarily a function of the steady state
aerodynamics that is discussed. The analysis presented provides a method for determining average
loads on a wind turbine. However, a number of important steady state and dynamic effects that cause
increased loads or decreased power production from those expected with the BEM theory presented
here, especially increased transient loads. Figure 3.7 gives the obtained result of the force analysis
on the airfoil sections. Figure 3.8 (a-d) show static pressure, velocity magnitudes, deformation and
stress distribution of the Horizontal-Axis Wind Turbine (HAWT).
3.2.1.1.11 Conclusion
In this study, we applied the finite element model of aerodynamics and static structural analyses of
Horizontal Axis Wind Turbine (HAWT). The first part of the paper focused on the wind turbine
geometry modeling, mesh generation, and numerical simulation of Horizontal-Axis Wind Turbine
(HAWT). The fluid and structural meshes are compatible at the interface and may be employed for
the coupled FSI analysis. These aerodynamic models have been coupled with a nonlinear formulation
describing the structural dynamics to moderate deformations. A comprehensive look at blade design
has shown that an efficient blade shape is defined by aerodynamic calculations based on chosen
parameters and the performance of the selected airfoils. Aesthetics plays only a minor role. The
optimum efficient shape is complex consisting of airfoils sections of increasing width, thickness and
twist angle towards the hub. This general shape is constrained by physical laws and is unlikely to
change. However, airfoils lift and drag performance will determine exact angles of twist and chord
lengths for optimum aerodynamic performance. Due to the large and flexible structure of the wind
turbine blades, there will probably be aeroelastic instability. The displacement of the tip of the blade
at the nominal wind speed (12 m/s) is obtained (0.045 m) a reduction in the power performance of
the turbine, which implies a reduction in the rated power. In order to do the intensive study of the
structural models and aeroelastic behavior of the blade, the aerodynamic is constructed correctly.
3.2.2 Flow in Multiple Rotating Reference Frames (MRF)
Many problems involve multiple moving parts or contain stationary surfaces which are not surfaces
of revolution (and therefore cannot be used with the Single Reference Frame modeling approach).
For these problems, you must break up the model into multiple fluid/solid cell zones, with interface
boundaries separating the zones. Zones which contain the moving components can then be solved
using the moving reference frame equations, whereas stationary zones can be solved with the
stationary frame equations. The manner in which the equations are treated at the interface lead to
two approaches:
➢ Multiple Reference Frame Model (MRF)
➢ Mixing Plane Model (MPM)
➢ Sliding Mesh Model (SMM)
Both the MRF and Mixing plane approaches are steady-state approximations, and differ primarily
in the manner in which conditions at the interfaces are treated. These approaches will be discussed
in following sections. The sliding mesh model approach is, on the other hand, inherently unsteady
due to the motion of the mesh with time. The MRF model is, perhaps, the simplest of the two
approaches for multiple zones. It is a steady-state approximation in which individual cell zones move
at different rotational and/or translational speeds. The flow in each moving cell zone is solved using
the moving reference frame equations. If the zone is stationary (ω = 0), the stationary equations are
used. At the interfaces between cell zones, a local reference frame transformation is performed to
enable flow variables in one zone to be used to calculate fluxes at the boundary of the adjacent zone.
It should be noted that the MRF approach does not account for the relative motion of a moving zone
48

with respect to adjacent zones (which may be moving or stationary); the grid remains fixed for the
computation. This is analogous to freezing the motion of the moving part in a specific position and
observing the instantaneous flow field with the rotor in that position. Hence, the MRF is often
referred to as the frozen rotor approach. While the
MRF approach is clearly an approximation, it can
provide a reasonable model of the flow for many
applications. For example, the MRF model can be
used for turbomachinery applications in which
rotor-stator interaction is relatively weak, and the
flow is relatively uncomplicated at the interface
between the moving and stationary zones. In
mixing tanks, for example, since the impeller-baffle
interactions are relatively weak, large-scale
transient effects are not present and the MRF
model can be used. Another potential use of the
MRF model is to compute a flow field that can be
used as an initial condition for a transient sliding
mesh calculation. This eliminates the need for a
startup calculation. The multiple reference frame Figure 3.9 Mixing Tank Geometry with One
model should not be used, however, if it is Rotating Impeller
necessary to actually simulate the transients that
may occur in strong rotor-stator interactions, the
sliding mesh model alone should be used.
3.2.2.1 Case Study – Mixing Tank
In a mixing tank with a single impeller, you can define a rotating reference frame that encompasses
the impeller and the flow surrounding it, and use a stationary frame for the flow outside the impeller
region. An example of this configuration is illustrated in Figure 3.9. (The dashes denote the interface
between the two reference frames). Steady-state flow conditions are assumed at the interface
between the two
reference frames. That is,
the velocity at the
interface must be the
same (in absolute terms)
for each reference frame.
The grid does not move.
You can also model a
problem that includes
more than one rotating
reference frame. Figure
3.10 shows a geometry
that contains two
rotating impellers side by
side. This problem would
be modeled using three Figure 3.10 Mixing Tank with Two Rotating Impellers
reference frames: the
stationary frame outside
both impeller regions and two separate rotating reference frames for the two impellers. (As noted
above, the dashes denote the interfaces between reference frames.)
49

3.2.3 The MRF Interface Formulation


The MRF formulation that is applied to the interfaces will depend on the velocity formulation being
used. The specific approaches will be discussed below for each case. It should be noted that the
interface treatment applies to the velocity and velocity gradients, since these vector quantities
change with a change in reference frame. Scalar quantities, such as temperature, pressure, density,
turbulent kinetic energy, etc., do not require any special treatment, and thus are passed locally
without any change.
3.2.3.1 Interface Treatment: Relative Velocity Formulation
In implementation of the MRF model, the calculation domain is divided into subdomains, each of
which may be rotating and/or translating with respect to the laboratory (inertial) frame. The
governing equations in each subdomain are written with respect to that subdomain's reference
frame. Thus, the flow in stationary and translating subdomains is governed by Continuity and
Momentum Equations, while the flow in
rotating subdomains is governed by the
equations presented in Equations for a
Rotating Reference Frame. At the
boundary between two subdomains, the
diffusion and other terms in the governing
equations in one subdomain require
values for the velocities in the adjacent
subdomain (see Figure 3.11). To enforce
the continuity of the absolute velocity, u,
is to provide the correct neighbor values
of velocity for the subdomain under
consideration. (This approach differs
from the mixing plane approach described
previously; The Mixing Plane Model,
where a circumferential averaging
technique is used). When the relative
velocity formulation is used, velocities in
each subdomain are computed relative to Figure 3.11 Interface Treatment for the MRF Model
the motion of the subdomain. Velocities
and their gradients are converted from a moving reference frame to the absolute inertial frame using
following equation. For a translational velocity ut, we have

𝐮 = 𝐮f + 𝛚
⏟×𝐫 + 𝐮t , ∇𝐮 = ∇𝐮f + ∇(𝛚 × 𝐫)
Swirl Velocity
Eq. 3.6
Note that scalar quantities such as density, static pressure, static temperature, species mass fractions,
etc., are simply obtained locally from adjacent cells.
3.2.3.2 Interface Treatment: Absolute Velocity Formulation
When the absolute velocity formulation is used, the governing equations in each subdomain are
written with respect to that subdomain's reference frame, but the velocities are stored in the absolute
frame. Therefore, no special transformation is required at the interface between two subdomains.
Again, scalar quantities are determined locally from adjacent cells.
50

3.2.3.3 Case Study 1 - Experiments and CFD Calculations on the Performance of a Non-Reversible
Axial Fan36

Authors : BACAK Aykut, ÜNVERDİ, Salih Özen


Conference : Proceedings of the 7th International Conference on Heating, Ventilation and Air
Conditioning, May 30-1st June, 2016, Tehran, Iran ICHVAC7-1813
Here, the CFD simulation results of a six blade axial fan are compared with test data obtained from
AMCA chamber. Flow through one period of the axial fan is simulated. In flow simulations standard
k-ε turbulence model with standard wall functions are implemented. Convergence of the simulation
results are tested under grid refinement by generating coarse, medium and fine meshes. Test results
are obtained from a test rig, designed according to AMCA-210 Standard.
3.2.3.3.1 Introduction
Axial fans are continuous flow machines that
provide positive or negative pressure for a space
by rotation of axial fan blades. Air flow is in the
axial direction. Axial fans are widely used at
parking garages, buildings, public areas such as
restaurants, hospitals and schools to provide fresh
air or expel the polluted air. Axial fans are classified
as propeller axial fans, vane axial fans and tube
axial fans. Various types of axial fans are used for
different purposes such as providing required air
flow rate or static pressure increase. An industrial
axial fan produced by Cvsair, Turkey is shown in
Figure 3.12.
Axial fans’ performance is the key factor that
determines the fire scenario in parking garages. In
parking garages, fresh or exhaust air flow rate is Figure 3.12 An Industrial Axial Fan Used For
determined according to the volume of the lot and Ventilation
required air exchange rate. Multiplying the parking
garage volume with air change rate gives the
exhaust flow rate and 50-70% of this exhaust flow rate is generally suitable as fresh air flow rate.
Therefore, performance characteristics of a fan used in a parking garage is the key factor for
ventilation.
Static pressure increase of axial fans is important when using duct systems or axial flow shafts.
Because continuous and local head losses are determined by the duct geometry and air flow rate, the
diameters and number of embranchments of an air flow shaft must be known to determine the
required air pressure rise provided by fans.
CFD analysis of fans with forward and backward skewed blade profiles are carried out by the
standard or RNG k-ε turbulence models and some of them are compared with test results before37-38.
Aerodynamic and aero-acoustic performance of an axial flow fan is optimized for efficiency and
pressure rise and sound pressure level by solving Reynolds averaged Navier-Stokes equations
coupled with shear stress transport turbulence model and Flows Williams-Hawkings equations and

36 BACAK Aykut, ÜNVERDİ, Salih Özen, “Experiments and CFD Calculations on the Performance of a Non-
Reversible Axial Fan”, Proceedings of the 7th International Conference on Heating, Ventilation and Air-
Conditioning, May 30-1st June, 2016, Tehran, Iran ICHVAC7-1813.
37 D. Dwivedi, D.S. Dantodiya, “CFD Analysis of Axial Flow Fans with Skewed Blades”, IJETAE, Issue 10, 2013).
38 T. Köktürk, “Design and Performance Analysis of a Reversible Axial Flow Fan”, Master Thesis, METU, 2005.
51

implementing Latin Hypercube sampling in the design space and multi objective evolutionary
algorithm coupled with response surface approximation surrogate model39-40.
3.2.3.3.2 Experimental Method
Sufficient ventilation of a parking garage depends on the correct fan selection. Fans are selected by
deciding if the air flow rate determined by intersecting the fans’ H-Q characteristic curves with the
system resistance curve is sufficient for the ventilation of the space. For that purpose, database of
various fans’ performance curves and static pressure loss vs. airflow rate curve of the ventilated
system are needed. Fan databases are setup by testing the performance of fans in test systems
that are designed according to the national and/or international standards. One of the commonly
used measurement systems is AMCA 210 test rig shown in Figure 3.13. The flow rate through the
fan is changed by adjusting the flow resistance, i.e. nozzle diameter, which is located in the variable
exhaust system of the AMCA chamber.

Figure 3.13 AMCA 210 test rig [8]

3.2.3.3.3 Numerical Method


The CFD solution domain consists of one slice of the axial fan around a blade, which is extended in
both upstream and downstream directions to improve the solution accuracy, as described below.
After CAD geometry cleanup and generation of surface and volume mesh required by the Multiple
Reference Frame, MRF, method, the grid is read into the computational fluid dynamics (CFD)
software. Then, the fluid properties, turbulence model, wall functions, fan rpm and boundary
conditions are set. Results are obtained by a parallel CFD solver and post-processed to get visual and
numeric data about the flow field.
The solution of a case is used as the initial guess for the next one, thereby a parametric study of the
fan performance is realized. According to the standards on the methods of fan testing41 the minimum

39 J.H. Kim, J.W. Kim, K.Y. Kim, “Axial-Flow Ventilation Fan Design Through Multi-Objective Optimization to
Enhance Aerodynamic Performance”, Journal of Fluids Engineering, vol. 133 / 101101-1, 2011.
40 J. Kim, B. Ovgor, K. Cha, J. Kim. S. Lee, K. Kim, “Optimization of the Aerodynamic and Aero acoustic Performance

of an Axial-Flow Fan”, AIAA Journal, Vol. 52, No. 9, 2014.


41 AMCA 210: Air Movement and Control Association International, Inc. “Laboratory Methods of Testing Fans for

Aerodynamic Performance Rating “, 1999.


52

length for both the upstream and downstream ducts of the axial fan is assumed to be 5D, where D is
the diameter of the fan casing. Both stationary and rotating zones of the MRF model are generated
for one sixth of the axial fan by using periodic boundary conditions. Velocity inlet and pressure outlet
boundary conditions are used at the upstream and downstream ends of the computational domain,
respectively.
[Pascu] built a CFD model of an axial fan which is
3D and 2D long in the upstream and downstream
sides of the fan, respectively42. The complete
surface mesh of the rotating zone obtained by
revolving one slice of the blade and the hub
around the fan axis is shown in Figure 3.14.
Surface mesh is created on both pressure and
suction sides of the fan blades, therefore the
blade thickness is maintained. The cylindrical
outer wall of the AMCA chamber is the boundary
that surrounds the computational model.
Boundary layer mesh is applied throughout
stationary and rotating zone walls and around
the fan blade walls.
There is not a stator in the model. Ten boundary
mesh layers are generated next to the duct wall
and five boundary mesh layers are used around Figure 3.14 Surface mesh of the rotating zone
the blades. Thin boundary mesh layers required
by the k-ε turbulence model are used to resolve the small blade tip clearance and flow around the
blades, which yield accurate prediction of the fan performance and especially the fan noise
generation43-44. Convergence of the results are
tested under grid refinement by generating three Mesh Density Number of Cells
meshes of different resolution, whose number of Coarse Mesh 169477
cells are given in Table 3.1.
Medium Mesh 178410
Reynolds averaged continuity and Navier-Stokes
Fine Mesh 496616
Equations coupled with turbulent kinetic energy
and dissipation rate transport equations of the
Table 3.1 Mesh Resolution and Number of Cells
standard k-ε turbulence model are used in the fan
simulations. RNG k-ε turbulence model chosen in
a similar study of axial flow fans 45. Continuity, momentum and k-ε turbulence model equations are
given below:
∂ρ ∂(ρui )
+ =0
∂t ∂xi
∂ ∂
ρui +
̅̅̅̅ ρu
̅̅̅̅̅
u =
∂t ∂xi i j

42 M. T. Pascu, “Modern Layout and Design Strategy for Axial Fans”, Ph.D. Thesis at Erlangen University, 2009.
43 Srinivas G., Srinivasa Rao Potti: “Numerical Simulation of Axial Flow Fan Using Gambit and Fluent”. IJRET,
Vol.3, No 3, pp. 586-590, 2014.
44 A. Raj. S, P. Pandian P.: “Effect of Tip Injection on an Axial Flow Fan under Distorted Inflow”. IJASER, Vol.3, No

10, pp.302-309, 2014.


45 T. Köktürk, “Design and Performance Analysis of a Reversible Axial Flow Fan”, Master Thesis, METU, 2005.
53

∂p ∂ ∂ui ∂uj 2 ∂ui ∂


− + [μ ( + − δij )] + ̅̅̅̅̅
(−ρu ′ u′ )
i j
∂xi ∂xj ∂xj ∂xj 3 ∂xi ∂xj
Eq. 3.7

∂κ ∂κ ∂κ ∂κ
ρ
+ ρU + ρV + ρW =
∂t ∂x ∂y ∂z
∂ μt ∂κ ∂ μt ∂κ ∂ μt ∂κ
( )+ ( )+ ( ) − ρε +
∂x σκ ∂x ∂y σκ ∂y ∂z σκ ∂z
∂U 2 ∂V 2 ∂W 2 ∂U ∂V 2 ∂U ∂W 2 ∂U ∂W 2
μt {2 [( ) + ( ) + ( ) ] + ( + ) + ( + ) +( + ) }
⏟ ∂x ∂y ∂z ∂y ∂x ∂z ∂x ∂z ∂x
𝐺
Eq. 3.8
∂ε ∂ε ∂ε ∂ε
ρ
+ ρU + ρV + ρW =
∂t ∂x ∂y ∂z
∂ μt ∂ε ∂ μt ∂ε ∂ μt ∂ε ε2 ε
( )+ ( )+ ( ) − C2 ρ + C1 μt G
∂x σκ ∂x ∂y σκ ∂y ∂z σκ ∂z κ κ
Eq. 3.9
After generating
tetrahedral cells with
appropriate resolution
above the inflation layers,
boundary conditions are
set. In Figure 3.15 and
Table 3.2 rotating and
stationary cell zones and
boundary conditions are
indicated. For more
accurate solutions, second
order discretization is
used for continuity,
momentum and Figure 3.15 Cell zones and boundary conditions
turbulence model
equations.
Turbulence Modeling κ-ε
3.2.3.3.4 Results and Inlet B.C. Velocity Inlet
Discussion Outlet B.C. Pressure Outlet
Reynolds averaged
Blade Pressure Side Periodic
continuity, Navier-Stokes
Blade Suction Side Periodic
and k-ε turbulence model
equations are solved Fluid Air
simultaneously using Fan Speed 1500
Multiple Reference Residual For Iteration Convergence 1E-5
Frames, MRF, method. Near Wall Treatment Standard Wall Function
After simulations, validity
of the results are checked Table 3.2 Solution Methods and Boundary Conditions
54

by verifying that y+ values on the pressure and suction surfaces of the fan blades remain between 30
and 100.
In the following figures comparison of CFD simulation and test results for fan pressure rise, torque
and efficiency vs. volumetric flow rate curves are given for the three grid resolutions. Difference
between the calculated and test torque data is caused by neglecting mechanical friction in the CFD
model.
30 ∗ Q ∗ ∆p
η=
π∗n∗T
Eq. 3.10
where Q is the volumetric airflow rate
in m3/s, ΔP is the fan pressure rise in
Pa, n is the fan speed in rpm, and T is
the sum of the fan aerodynamic and
mechanical friction torques in Nm. In
Figure 3.17, numerical torque values
are almost half of the experimental
ones because the losses caused by the
bearing and transmission friction
torques are ignored. Even though
there are significant variations in the
previous charts, simulation results of
fan static pressure rise vs. volume flow
rate curves are quite similar to the test
result, because mechanical friction of
bearings and the transmission do not
affect these parameters. Following
formula is used to calculate the electric
motor torque in terms of the motor Figure 3.17 Torque vs. Volumetric Flow Rate Curves
power and fan rpm.

60 ∗ p
Tmotor =
2∗π∗n
Eq. 3.11
The torque curve which is obtained by
adding the aerodynamic torque
calculated by postprocessing
CFD results and the motor torque
calculated above and estimated
mechanical friction torque is compared
with the torque value read from electric
control panel. Figure 3.16 shows the
streamline inside the duck at the rotor.
3.2.3.3.5 Conclusion
CFD simulation results of an axial fan are
obtained under grid refinement and are
compared with the fan test data. Use of
the CFD model of a slice of the fan with Figure 3.16 The streamlines inside the duct at a rotor
speed of 1500 rpm
periodic boundary conditions decreases
55

the memory requirements of the simulations because of the substantial decrease of the number of
cells. Acceptable agreement of the axial fan static pressure rise vs. airflow rate curve with the test
data is promising for using CFD as a means of predicting the performance of axial fans. The pressure
rise prediction of the current CFD model is quite well in the neighborhood of the design point of the
fan. Indeed, the calculated pressure rise corresponding to the volume flow rate at the design point is
almost the same as the test data. As the fan operates away from the design point and stall inception
occurs, numerical results begin to move away from the test data. This shows that the standard k-ε
turbulence model with standard wall functions is insufficient in calculating the location of the
separation point and the pressure change on the blade surfaces for separated flows. If electric motor
and mechanical friction torques are not taken into account, aerodynamic torque calculated by CFD is
almost half of the value measured in experiments. As the motor torque and estimated mechanical
friction torque values are considered, numerical and experimental torque vs. air flow-rate curves get
closer. Furthermore, because the calculated efficiency depends on the friction torque, taking into
account of the total torque results in having closer agreement between the predicted and measured
efficiency values. Reversible axial fan performance and noise predictions by CFD simulations and
comparison of the results with the test data is planned as the future research. Furthermore the degree
of improvement of the predicted fan performance at stall condition, provided by applying shear
stress transport, SST, k-ω turbulence model on fine meshes is to be studied in the future.
3.2.3.4 Case Study 2 - Potsdam Propeller CFD Benchmark
Authors : Daniel LaCroix1 and Radek Máca2
Affiliations : 1Pointwise, 2CFD Support
Originally Appeared : LinkedIn – CFD Support
Source : Same as Above
This report presents the benchmark validation of CFD simulation results of the Potsdam Propeller
Test Case (PPTC), using TCFD® with POINTWISE® mesh. PPTC is a marine propulsor that was
extensively measured by SVA Potsdam and related data were published [1] [2] [3]. The aim of this
benchmark was to evaluate the TCFD® , computational fluid dynamics (CFD) software, on the very
advanced mesh, created in POINTWISE® , mesh generation software, and to compare the results with
the measurement data available. The particular goal of this benchmark is to compare the propeller
Efficiency, Torque Coefficient, and Thrust Coefficient vs. Advance Coefficient with the real
experimental measurement of SVA Potsdam Laboratory.
3.2.3.4.1 Potsdam Propeller CFD Benchmark Description
The high demand for improving the accuracy, quality, and credibility of the CFD simulation results,
should be assessed by providing a high qualitative and intensive comparison with experimental
measurement data. The purpose of this benchmark is the validation of CFD simulation software
TCFD® with the mesh created in high-end meshing software
POINTWISE® and to compare the results with the
measurement data available. Potsdam Propeller Test Case
(PPTC) is a marine propulsor that was extensively measured
by SVA Potsdam and related data were published in [1], [2],
and [3].
The particular goal of this benchmark is to compare the
propeller Efficiency, Torque Coefficient, and Thrust
Coefficient vs. Advance Coefficient with the real
experimental measurement of SVA Potsdam Laboratory. The
experimental investigation includes open water test and
Figure 3.18 SVA Potsdam
velocity field measurements at different operation
Laboratory
conditions. A detailed description of the open water tests
56

conducted at the towing tank of the SVA is presented in the SVA report [1], which can be found on
the SVA website (sva-potsdam.de). See Figure 3.18. At the propeller analysis, there are a few
important dimensionless numbers. Those are Advance Coefficient, Thrust Coefficient, Torque
Coefficient, and Efficiency. They are defined as (respectively):

Va T Q JK T
J= , KT = , KQ = , η0 =
n. D ρ. n2 . D4 ρ. n2 . D5 2πK Q
Eq. 3.12
Where Va is the advance speed [m/s], n is the speed of rotation [1/s], D is propeller diameter [m], T
is thrust [N], ρ is water density [kg/m3], Q is torque [Nm]. The measurement results are available
for Advance coefficients from J = 0.5 to 1.6 and the simulation points are chosen accordingly.
Altogether, 11 Advance Coefficient modes were simulated according to the measurements.
3.2.3.4.2 Meshing
An unstructured viscous computational mesh was constructed with Pointwise on the Potsdam
propeller geometry as part of this TCFD validation benchmark. Pointwise, Inc. has previously worked
with variants of the geometry for other studies. For an in-depth discussion of Pointwise technology
and how it can be used for this particular geometry, please consult [4].
A combination of anisotropic and isotropic triangles were used in the surface mesh discretization.
Areas of high curvature - such as the leading edge, trailing edge, and the tip were resolved by utilizing
Pointwise’s T-Rex algorithm. This tool grows anisotropically stretched, right-angled triangles layered
in the normal direction to a boundary [5], as shown below. Using this, areas of high curvature are
able to be resolved without the need to isotopically refine the area. The result is an accurate
adherence to the surface and a reduction in the point count. The interior of the surface mesh was
resolved with isotropic triangles created using a modified Delaunay algorithm. (see Figure 3.19).

Figure 3.19 Potsdam Propeller Benchmark - POINTWISE Mesh


57

After meshing the geometry, the outer boundary of the moving reference frame (MRF) as well as the
outer boundary of the computational domain were meshed utilizing isotropic triangles and the
Delaunay algorithm. The MRF is cylindrical domain approximately 4.8 D long (4.8 x the propeller
diameter) and 1.5 D in diameter (see Figure 3.20). It starts just upstream of the propeller and
extends downstream into the wake. A far field block was generated outside of the MRF
corresponding to 10 D and 2.6 D; these were the limits taken from the file provided.

Figure 3.20 Domain Decomposition of MRF

The volume mesh is a combination of a prismatic core surrounded by isotropic tetrahedral cells. The
prismatic portions of the grid were initialized using T-Rex. Starting from the surface mesh,
anisotropic tetrahedral cells were grown until reaching a desired stop criteria, colliding with another

Figure 3.21 Potsdam Propeller Benchmark - snappyHexMesh Mesh


58

front, or violating quality criteria. If an element stops advancing this did not prevent adjacent
elements from continuing. After the tetrahedral layers are grown, the cells are combined to form
prisms (or hexagons if the surface mesh is made up of quadrilateral cells). This reduces the total cell
count of the mesh without sacrificing quality. Once the near-wall viscous mesh was generated, the
remainder of the volume was populated with isotropic tetrahedral cells. The total cell count was just
below 4.1 M46 cells. The average maximum included angle was 101, and the maximum was 170. The
average volume ratio was 1.8 with a maximum of 28.
3.2.3.4.3 TCFD® Simulation Setup
The simulation run in TCFD is quite straightforward. The external mesh, created in POINTWISE, is
simply loaded and the simulation parameters are set. The simulation type is the propeller. Time
management is steady-state. The fluid flow model is incompressible. The mesh has two components
(the water tunnel and the cylinder with propeller inside) of total 4.1M cells. The inlet flow velocity
defines the advance coefficient and the velocity varies from 1.7 to 5.9 m/s in 11 points. The outlet
boundary condition is static pressure. For turbulence modeling, the RANS modeling approach has
been used with the k-ω SST turbulence model with the wall functions. The fluid properties of water
are selected. The density of 997.71 kg/m3. Dynamic viscosity of 9.559e-4 Pa.s.
3.2.3.4.4 TCFD® Simulation Post-processing
TCFD® includes a built-in post-processing module that automatically evaluates the required total
quantities, such as propeller efficiency, thrust coefficient, torque coefficient, forces, force coefficients,
flow rates, and much more. All these quantities are evaluated throughout the simulation run, and all
the important data is summarized in tabled .csv files as well as in the HTML report, which can be
updated anytime during the simulation for every run. Furthermore, visual postprocessing of the
volume fields can be done with ParaView. (see Figure 3.22).

Figure 3.22 Surface Static Pressure

46 M= million
59

3.2.3.4.5 Efficiency, Thrust & Torque Coefficient vs. Advance Coefficient


A propeller is the most common propulsor on ships, imparting momentum to a fluid which causes a
force to act on the ship. The most comprehensive propeller performance information is provided by
displaying all the key characteristics into one graph. See Figure 3.23.

Figure 3.23 Comparison of TCFD vs. SVA using Pointwise Mesh

3.2.3.4.6 Conclusion
The CFD analysis of the PPTC was performed successfully. It has been shown that the TCFD® in
connection with POINTWISE® provides very accurate results that are in perfect agreement with the
measurement data. All the simulation and measurement data are freely available. Potential questions
about TCFD® are to be sent to info@cfdsupport.com . Questions about POINTWISE® are to be sent to
pointwise@pointwise.com .
3.2.3.4.7 References
[1] Barkmann, U., Potsdam Propeller Test Case (PPTC) - Open Water Tests with the Model Propeller
VP1304, Report 3752, SchiffbauVersuchsanstalt Potsdam, April 2011
[2] Barkmann, U., Heinke, H.-J., Potsdam Propeller Test Case (PPTC) Test Case Description, Second
International Symposium on Marine Propulsors smp’11, Hamburg, Germany, June 2011, Workshop:
Propeller performance
[3] Heinke, H.-J., Potsdam Propeller Test Case (PPTC), Cavitation Tests with the Model Propeller
VP1304, Report 3753, SchiffbauVersuchsanstalt Potsdam, April 2011
[4] Carrigan, T., Bagheri, B., “A Study of the Influence of Meshing Strategies on CFD Simulation
Efficiency,” NAFEMS World Congress 2017, NWC17-466, 2017.
[5] Steinbrenner, J. P. and Abelanet, J.P., "Anisotropic Tetrahedral Meshing Based on Surface
Deformation Techniques," AIAA-20060554, AIAA 45th Aerospace Sciences Meeting, Reno, NV
60

3.3 The Mixing Plane Model (MPM)


The mixing plane model provides an alternative to the multiple reference frame and sliding mesh
models for simulating flow through domains with one or more regions in relative motion. The
Multiple Reference Frame Model, is applicable when the flow at the boundary between adjacent
zones that move at different speeds is nearly uniform (mixed out). If the flow at this boundary is not
uniform, the MRF model may not provide a physically meaningful solution. The sliding mesh model
(Sliding Mesh Theory) may be appropriate for such cases, but in many situations it is not practical to
employ a sliding mesh. For example, in a multistage turbomachine, if the number of blades is
different for each blade row, a large number of blade passages is required in order to maintain
circumferential periodicity. Moreover, sliding mesh calculations are necessarily unsteady, and thus
require significantly more computation to achieve a final, time-periodic solution. For situations
where using the sliding mesh model is not feasible, the mixing plane model can be a cost-effective
alternative. In the mixing plane approach, each fluid zone is treated as a steady-state problem. Flow-
field data from adjacent zones are passed as boundary conditions that are spatially averaged or mixed
at the mixing plane interface. This mixing removes any unsteadiness that would arise due to
circumferential variations in the passage-to-passage flow field (e.g., wakes, shock waves, separated
flow), thus yielding a steady-state result. Despite the simplifications inherent in the mixing plane
model, the resulting solutions can provide reasonable approximations of the time-averaged flow
field.
3.3.1 Rotor and Stator Domains
Consider the turbomachine stages shown schematically in Figure 3.24 and Figure 3.25. In each
case, the stage consists of two flow domains: the rotor domain, which is rotating at a prescribed
angular velocity, followed by the stator domain, which is stationary. The order of the rotor and stator
is arbitrary (that is, a situation where the rotor is downstream of the stator is equally valid). In a
numerical simulation, each domain will be represented by a separate mesh. The flow information
between these domains will be coupled at the mixing plane interface (as shown in Figure 3.25 and
Figure 3.24) using the mixing plane model. Note that you may couple any number of fluid zones in
this manner; for example, four blade passages can be coupled using three mixing planes. Note that
the stator and rotor meshes do not have to be conformal; that is, the nodes on the stator exit

Figure 3.24 Mixing Plane Concepts as Applied to Axial Rotation


61

boundary do not have to match the nodes on


the rotor inlet boundary. In addition, the
meshes can be of different types (e.g., the
stator can have a hexahedral mesh while the
rotor has a tetrahedral mesh).
3.3.2 The Mixing Plane Concept
The essential idea behind the mixing plane
concept is that each fluid zone is solved as
a steady-state problem. At some
prescribed iteration interval, the flow data
at the mixing plane interface are averaged in
the circumferential direction on both the
stator outlet and the rotor inlet boundaries.
By performing circumferential averages at
specified radial or axial stations, profile of
flow properties can be defined. These
profiles which will be functions of either the
axial or the radial coordinate, depending on Figure 3.25 Mixing Plane Concepts Applied to Radial
the orientation of the mixing plane are then Rotation
used to update boundary conditions along
the two zones of the mixing plane interface. In the examples shown in Figure 3.25 and Figure 3.24,
profiles of averaged total pressure (p0), direction cosines of the local flow angles in the Radial,
Tangential, and Axial directions (αr; αt; αz), total temperature (T0), turbulence kinetic energy (k),
and turbulence dissipation rate (ε) are computed at the
rotor exit and used to update boundary conditions at the Up Stream Down Stream
stator inlet. Likewise, a profile of static pressure (ps), Pressure Outlet Pressure Inlet
direction cosines of the local flow angles in the radial, Pressure Outlet Velocity Inlet
tangential, and axial directions (αr ; αt ; αz), are computed Pressure Outlet Mass flow Inlet
at the stator inlet and used as a boundary condition on the
rotor exit. Passing profiles in the manner described above Table 3.3 Prescribed Boundary zone
assumes specific boundary condition types have been for Mixing Plane
defined at the mixing plane interface. The coupling of an
upstream outlet boundary zone with a downstream inlet boundary zone is called a mixing plane
pair. In order to create mixing plane pairs, the boundary zones must be as prescribed as
Table 3.3.
3.3.3 Mixing Plane Algorithm
The basic mixing plane algorithm can be described as follows:
• Update the flow field solutions in the stator and rotor domains.
• Average the flow properties at the stator exit and rotor inlet boundaries, obtaining profiles
for use in updating boundary conditions.
• Pass the profiles to the boundary condition inputs required for the stator exit and rotor inlet.
• Repeat steps 1-3 until convergence.

Note that it may be desirable to under-relax the changes in boundary condition values in order to
prevent divergence of the solution (especially early in the computation).
3.3.3.1 Mass Conservation Across the Mixing Plane
Note that the algorithm described above will not rigorously conserve mass flow across the mixing
plane if it is represented by a pressure inlet and pressure outlet mixing plane pair. If you use a mass
62

flow inlet and pressure outlet pair instead, we will force mass conservation across the mixing plane.
The basic technique consists of computing the mass flow rate across the upstream zone (pressure
outlet) and adjusting the mass flux profile applied at the mass flow inlet such that the downstream
mass flow matches the upstream mass ow. This adjustment occurs at every iteration, thus ensuring
rigorous conservation of mass ow throughout the course of the calculation. Also note that, since mass
flow is being fixed in this case, there will be a jump in total pressure across the mixing plane. The
magnitude of this jump is usually small compared with total pressure variations elsewhere in the
flow field. Other quantities which will be conserved across Mixing Plane are Swirl and Total
Enthalpy.

3.4 Sliding Mesh Modeling


In sliding meshes, the relative
motion of stationary and rotating
components in a rotating
machine will give rise to
unsteady interactions. These
interactions are illustrated in
Figure 3.26, and generally
classified as follows:
• Potential interactions:
flow unsteadiness due to
pressure waves which
propagate both upstream
and downstream. Figure 3.26 Illustration of Unsteady Interactions
• Wake interactions: flow
unsteadiness due to wakes from upstream blade rows, convecting downstream.
• Shock interactions: for transonic/supersonic ow unsteadiness due to shock waves striking
the downstream blade row.
Where the multiple reference frame (MRF) and mixing plane (MP) models, are models that are
applied to steady-state cases, thus neglecting unsteady interactions, the sliding mesh model cannot
neglect unsteady interactions. The sliding mesh model accounts for the relative motion of stationary
and rotating components.
3.4.1 Sliding Mesh Theory
When a time-accurate solution for rotor-stator interaction (rather than a time-averaged solution) is
desired, you must use the sliding mesh model to compute the unsteady flow field. As mentioned in
Section 10.1: Introduction, the sliding mesh model is the most accurate method for simulating flow
in multiple moving reference frames, but also the most computationally demanding. Most often, the
unsteady solution that is sought in a sliding mesh simulation is time periodic. That is, the unsteady
solution repeats with a period related to the speeds of the moving domains. However, you can model
other types of transients, including translating, sliding mesh zones (e.g., two cars or trains passing in
a tunnel). Reminder that for flow situations where there is no interaction between stationary and
moving parts (i.e., when there is only a rotor), the computational domain can be made stationary by
using a rotating reference frame. (ω = 0). When transient rotor-stator interaction is desired (as in
the examples in Figure 3.27 (a) and Figure 3.27 (b), you must use sliding meshes. If you are
interested in a steady approximation of the interaction, you may use the multiple reference frame
model or the mixing plane model, as described before.
63

(a) Rotor-Stator
Interaction (b) Blower

Figure 3.27 Examples of Transient Interaction using Sliding Mesh

3.4.2 The Sliding Mesh Technique


In the sliding mesh technique two or more cell zones are used. (If you generate the mesh in each zone
independently, you will need to merge the mesh profiles prior to starting the calculation. Each cell
zone is bounded by at least one interface zone where it meets the opposing cell zone. The interface
zones of adjacent cell zones are associated with one another to form a grid interface. The two cell
zones will move relative to each other along the grid interface. Be advised that the grid interface must
be positioned so that it has fluid cells on both sides. For example, the grid interface for the geometry
shown in Figure 3.28-(c) must lie in the fluid region between the rotor and stator; it cannot be on
the edge of any part of the rotor or stator. During the calculation, the cell zones slide (i.e., rotate or
translate) relative to one another along the grid interface in discrete steps. To recap, topological Mesh
Changes in Sliding Interface are:
• Defined by a master and slave surfaces.
• As surfaces move relative to each other, perform mesh cutting operations and replace original
faces with facets.
• Re-assemble mesh connectivity on all cells and faces touching the sliding surface: fully
connected 3-D mesh.
• Once the mesh is complete, there is no further impact!
• Connectivity across interface changes with relative motion (see Figure 3.28 (a-d)).
Figure 3.28-(a) and Figure 3.28-(b) show the initial position of two grids and their positions after
some translation has occurred. For an axial rotor/stator configuration, in which the rotating and
stationary parts are aligned axially instead of being concentric (see Figure 3.28-(d)), the interface
will be a planar sector. This planar sector is a cross-section of the domain perpendicular to the axis
of rotation at a position along the axis between the rotor and the stator.
3.4.3 Sliding Mesh Concept
As discussed before, the sliding mesh model allows adjacent grids to slide relative to one another. In
doing so, the grid faces do not need to be aligned on the grid interface. This situation requires a means
of computing the flux across the two non-conformal interface zones of each grid interface. To
compute the interface flux, the intersection between the interface zones is determined at each new
time step. The resulting intersection produces one interior zone (a zone with fluid cells on both sides)
and one or more periodic zones. If the problem is not periodic, the intersection produces one interior
zone and a pair of wall zones (which will be empty if the two interface zones intersect entirely), as
shown in Figure 3.28 (a). (You will need to change these wall zones to some other appropriate
64

boundary type.) The resultant interior zone corresponds to where the two interface zones overlap;
the resultant periodic zone corresponds to where they do not. The number of faces in these
intersection zones will vary as the interface zones move relative to one another. Principally, fluxes
across the grid interface are computed using the faces resulting from the intersection of the two
interface zones, rather than from the interface zone faces themselves.

(b) Sliding Mesh


(a) Initial Position

(c) Rotor /Starter


Interactions (d) Linear Grid Interface

Figure 3.28 Initial position and some translation with Sliding Interface

Figure 3.28
In the example shown in Figure 3.28 (b), the interface zones are composed of faces A-B and B-C,
and faces D-E and E-F. The intersection of these zones produces the faces a-d, d-b, b-e, etc. Faces
produced in the region where the two cell zones overlap (d-b, b-e, and e-c) are grouped to form an
interior zone, while the remaining faces (a-d and c-f) are paired up to form a periodic zone. To
compute the flux across the interface into cell IV, for example, face D-E is ignored and faces d-b and
b-e are used instead, bringing information into cell IV from cells I and III, respectively.
65

4 Elements of Turbomachinery
4.1 Background
Turbomachinery is widely used equipment in industry such as compressors and turbines in a jet
engine; steam turbine in power plants, propeller for ships, hydraulic turbines for irrigation, wind
turbines for green energy, small fans for cooling, and so on (Wang, 2010)47. A common feature of
these devices is that they all work with fluid and have rotating component. Gorla48 gives a general
definition of turbomachinery which says “Turbomachinery is a device in which the energy transfer
occurs between a flowing fluid and a rotating element due to dynamic action, and results in a
change in pressure and momentum of the fluid”. The usage of turbomachinery has a long history.
It is recorded that the waterwheel, a kind of primitive turbomachinery, was invented and used for
power generation more than hundred years ago. Although the configuration is simple, it does follow
the same basic principle with other complicated modern turbomachinery’s, for instance the
compressor and the gas turbine in a jet engine. Figure 4.1 represents classification of
turbomachines. Here we concern mainly with axial devices. As the air is compressed in compressor
before entering the combustion chamber where it is mixed with fuel and combustion occurs (a.k.a.,
aggravated stage). Then the gas with high pressure and high temperature flows through gas turbines
and leaves the engine through a nozzle. While expanding through the turbine blades, power is
released from the gas and drives the turbine rotating. This constitutes the modern gas turbine engine

Axial Flow Devices

Flow Direction Centrifugal Flow Devices

Mixed Flow
Devices Wind Turbine
Turbomachines
Comprssible Gas Turbine

Steam Turbine
Fluid Physics

Pumps
Incompressible
Hydroulic Turbine

Figure 4.1 Classification of Turbomachines

47 Xiaodong Wang, “CFD Simulation of Complex Flows in Turbomachinery and Robust Optimization of Blade
Design”, Submitted to the Department of Mechanical Engineering Doctor of Philosophy, July 2010.
48 R. S. R. Gorla. Turbomachinery: Design and Theory. CRC Press, 2003.
66

phenomena to be discussed next.

4.2 Historical Perspectives


The gas turbine is an internal combustion (IC) engine that uses air as the working fluid. It is the
production of hot gas during fuel combustion, not the fuel itself that the gives gas turbines the name.
Gas turbines can utilize a variety of fuels, including natural gas, fuel oils, and synthetic fuels.
Combustion occurs continuously in gas turbines, as opposed to reciprocating IC engines, in which
combustion occurs intermittently. The engine extracts chemical energy from fuel and converts it to
mechanical energy using the gaseous energy of the working fluid (air) to drive the engine and
propeller, which, in turn, propel the airplane. The gas turbine engine was first invented in the
1930s∼1940s, which gave the opportunity of rapid development to turbomachinery. From the initial
turbojet engine to the modern turbofan engine with large bypass ratio, the evolution of jet engine
requires more advanced compressors and turbines with higher stage pressure ratio and higher
efficiency. Since 1988, the military of USA launched a series of research projects to develop advanced
turbines, such as “IHPTET” (Integrated High Performance Turbine Engine Technology), “VAATE”
(Versatile Affordable Advanced Turbine Engines) etc. The primary goal is to double the thrust-to-
weight ratio (TWR) of engine which will reach to 15∼20, decrease the fuel consumption ratio by 15%
∼ 30%. Compressor and turbine are two core components of jet engine. The performance of a jet
engine strongly depends on the design level of compressor and turbine. Therefore, significant
researching efforts have been spent on improving the performance of turbomachinery. Today, the
modern compressor stage has an efficiency of about 90% and the modern turbine stage has an
efficiency of up to 95%. Further improvements become more and more difficult and require much
deeper understanding of the flow field inside of the turbomachinery. Meanwhile, in industrial field,
steam turbine and gas turbine are the main instruments of power generation. Due to the energy
crisis, design of advanced turbine with higher efficiency is much more crucial than ever before.
Therefore, similar strong demands of improving the performance of turbomachinery are also
brought forward. While a turbine transfers energy from a fluid to a rotor, a compressor transfers
energy from a rotor to a fluid. These two types of machines are governed by the same basic
relationships including Newton's second Law of Motion and Euler's energy equation for
compressible fluids 49.

4.3 Modern Turbomachinery as Related to Gas Turbine Engine


In general, the rotating element is named rotor which is usually composed of one or several rows of
rotating blades. There also exits a stator which is also composed of rows of blades, but not rotating.
A pair of stator and rotor constitutes a stage. According to the way of energy transfer, turbomachines
are generally divided into two main categories. The first category is used primarily to generate power
which is called Turbine, including steam turbines, gas turbines and hydraulic turbines. The main
function of the second category is to increase the total pressure of the working fluid by consuming
power which includes compressors, pumps and fans as detailed in Figure 4.2. According to inlet
and outlet flow directions, turbomachines can be classified into two types: axial turbomachinery and
radial turbomachinery. However, here we concern only focuses on the axial turbomachinery. More
detail classification and description about the configurations can be found in50. Considerable
progress in development and application of CFD for aero-engines internal flow systems has been
made in recent years. CFD is regularly used in industry for assessment of air systems, and
performance of CFD for basic axisymmetric rotor/rotor and stator/rotor disc cavities with radial
through flow is largely understood and documented. In cooperation with 3D geometrical features
and calculation of unsteady flow are becoming common place. Automation of CFD, coupling with

49 See previous.
50 M. T. Schobeiri. Turbomachinery: Flow Physics and Dynamic Performance. Springer, Berlin, 2005. 2, 37, 38
67

thermal models of solid components, is current area of development. A wide variety of flow
phenomena, which are coupled in nature, occur in Turbomachinery CFD ranging from shock surfaces,
boundary layer, secondary flow, and vortex generating from blade tip and hob. These, makes the flow
analysis of turbo-machinery extremely complex and CFD limited. The number of turbine stages varies
in different types of engines, with high bypass ratio engines tending to have the most turbine stages.
The number of turbine stages can have a great effect on how the turbine blades are designed for each
stage. Many gas turbine engines are twin spool designs, meaning that there is a high pressure spool
and a low pressure spool. The high pressure turbine is exposed to the hottest, highest pressure air,
and the low pressure turbine is subjected to cooler, lower pressure air. The difference in conditions
leads to the design of high pressure and low pressure turbine blades that are significantly different
in material and cooling choices even though the aerodynamic and thermodynamic principles are the
same. Under these severe operating conditions inside the gas and steam turbines, the blades face high
temperature, high stresses, and potentially high vibrations. Steam turbine blades are critical
components in power plants which convert the linear motion of high temperature and high pressure
steam flowing down a pressure gradient into a rotary motion of the turbine shaft.
Figure 4.7 illustrates of a twin spool jet engine. The high pressure turbine (HPT) is connected by a
single spool to the high pressure compressor (purple) and the low pressure turbine is connected to
the low pressure compressor by a second spool (green)51.

Figure 4.2 Component of Turbomachines and their Thermodynamic (Brayton cycle) properties

4.4 Difference Among Turbojet, Turbofan and Turboprop Engines in Aviation


Both engines use a turbine for power. This is where the "turbo" part of the name comes from. In a
turbine engine, air is compressed and then fuel is ignited in this compressed air. The energy produced

51 From Wikipedia, the free encyclopedia.


68

by the ignition turns the turbine. The turbine is then able to drive both the compressor at the front
of the engine and also some useful load. In airplanes, it produces thrust.
4.4.1 Turbojet
The first jet engine was a turbojet. This is a simple turbine engine that produces all of its thrust from
the exhaust from the turbine section. However, because all of the air is passing through the whole
turbine, all of it must burn fuel. This means it is inefficient, and the solution is the turbofan (see
Figure 4.3).

Figure 4.3 Turbojet Engine

4.4.2 Turbofan
In a turbofan, the turbine primarily drives a fan at the front of the engine. Most engines drive the fan
directly from the turbine. There are usually at least two separate shafts to allow the fan to spin slower
than the inner core of the engine. The fan is surrounded by a cowl which guides the air to and from
the fan. Part of the air enters the turbine section of the engine, and the rest is bypassed around the
engine. In high-bypass engines, most of the air only goes through the fan and bypasses the rest of the
engine and providing most of the thrust (see Figure 4.4).

Figure 4.4 Turbofan Engine


69

4.4.3 Turboprop
In a turboprop, the turbine primarily drives a propeller at the front of the engine. There is no cover
around the prop. Some air enters the turbine, the rest does not. The propeller is geared to allow it to
spin slower than the turbine (see Figure 4.5). Although this diagram shows only a single shaft, many
turboprops have two, with a high pressure shaft driving the compressor and a low pressure shaft
driving the propeller. Turboprops are more efficient at lower speeds since the prop can move much
more air with a smaller turbine than the fan on a turbofan engine. The cover around the turbofan's
large fan allows it to perform better than an open propeller at high speeds, but limits the practical
size of the fan.

Figure 4.5 Turboprop Engine

At supersonic speeds, turbojets have more of a performance benefit. They develop all of their thrust
from the high velocity turbine exhaust, while turbofans supplement that with the lower velocity air
from the fan. Since the air from the fan is also not compressed nearly as much as the core turbine
flow, it is also harder to prevent the flow from going supersonic and causing losses. The Concorde
used turbojets because it was designed to cruise for long periods at supersonic speeds. Modern
fighter jet engines are turbofans, which
provide a compromise between
efficiency and speed.
Elsewhere in aviation, turbine engines
are used in helicopters, as a turboshaft
engine driving the rotors instead of a
propeller, and with a freewheeling
clutch to enable autorotation’s (see
Figure 4.6). Turbocharged piston
engines use a turbine much differently
from the examples above. Instead of
being the primary power source, the
turbine only assists the piston engine.
A turbocharger uses a turbine to
compress air sent to the engine intake.
The increased compression helps the Figure 4.6 Turboshaft Engine
engine generate more power. The
70

turbine of a turbocharger is driven by engine exhaust gasses, and a supercharger is similar but is
directly powered by the engine52.
4.4.4 How does it work?
Gas turbines are comprised of three primary sections mounted on the same shaft: the compressor,
the combustion chamber (or combustor) and the turbine, as described above. The compressor can
be either axial flow or centrifugal flow. Axial flow compressors are more common in power
generation because they have higher flow rates and efficiencies. Axial flow compressors are
comprised of multiple stages of rotating and stationary blades (or stators) through which air is drawn
in parallel to the axis of rotation and incrementally compressed as it passes through each stage. The
acceleration of the air through the rotating blades and diffusion by the stators increases the pressure
and reduces the volume of the air. Although no heat is added, the compression of the air also causes
the temperature to increase. The compressed air then mixed with fuel injected through nozzles. The
fuel and compressed air can be pre-mixed or the compressed air can be introduced directly into the
combustor. The fuel-air mixture ignites under constant pressure conditions and the hot combustion
products (what we like to call: aggravated gases) are directed through the turbine where it expands
rapidly and imparts rotation to the shaft. The turbine is also comprised of stages, each with a row of
stationary blades (or nozzles) to direct the expanding gases followed by a row of moving blades. The
rotation of the shaft drives the compressor to draw in and compress more air to sustain continuous
combustion. The remaining shaft power is used to drive a generator which produces electricity.
Approximately 55-65 % of the power produced by the turbine is used to drive the compressor. To
optimize the transfer of kinetic energy from the combustion gases to shaft rotation, gas turbines can
have multiple compressor and turbine stages. Because the compressor must reach a certain speed
before the combustion process is continuous – or self-sustaining – initial momentum is imparted to
the turbine rotor from an external motor, static frequency converter, or the generator itself. The
compressor must be smoothly accelerated and reach firing speed before fuel can be introduced and
ignition can occur. Turbine speeds vary widely by manufacturer and design, ranging from 2,000
revolutions per minute (rpm) to 10,000 rpm. Initial ignition occurs from one or more spark plugs
(depending on combustor design). Once the turbine reaches self-sustaining speed above 50% of full

Figure 4.7 Twin Pool Trubofan Jet Engine

52 Aviation weekly.
71

speed; the power output is enough to drive the compressor, combustion is continuous, and the starter
system can be disengaged.
Simply put, in a compressor, to raise the pressure, the fluid must be slowed down as it passes
through a blade row. In a turbine, to drop the pressure, the fluid must be accelerated as it passes
through a blade row. By having alternate stationary and moving blade rows and making use of
the change of frame of reference, it is possible to always slow down (relative to the blade row)
or always speed up the fluid. For example: In a turbine the flow is accelerated in the stator
(stationary blade row). However, because the rotor row is moving, the flow appears to be moving
more slowly in the relative frame and so can be re-accelerated in the relative frame. This
appears to be a deceleration in the absolute frame53.
4.4.5 What is Thrust?
Thrust is a mechanical force which is generated through the reaction of accelerating a mass of gas, as
explained by Newton's 3rd law of motion. A gas or working fluid is accelerated to the rear and the
engine and aircraft are accelerated in the opposite direction. To accelerate the gas, we need some
kind of propulsion system. For right now, let us just think of the propulsion system as some machine
which accelerates a gas. From Newton's second law of motion, we can define a force T to be the
change in momentum of an object with a change in time. The thrust force, using a simple control
volume around propulsion systems obtained (see Figure 4.8) as

Figure 4.8 A 1D Control Volume around a propulsion system (Courtesy’s of NASA Glen Research
Center)

if Pe ≠ P0 → T = ṁ e Ve − ṁ 0 V0 + (Pe − P0 )Ae
if Pe = P0 → T = ṁ e Ve − ṁ 0 V0
Where T is thrust and ṁ = mass flow rate
Eq. 4.1
We see that there are two possible ways to produce high thrust. One way is to make the engine flow
rate (m dot) as high as possible. As long as the exit velocity is greater than the free stream, entrance
velocity, a high engine flow will produce high thrust. This is the design theory behind propeller
aircraft and high-bypass turbofan engines. A large amount of air is processed each second, but the
velocity is not changed very much. The other way to produce high thrust is to make the exit velocity
very much greater than the incoming velocity. This is the design theory behind pure turbojets,
turbojets with afterburners, and rockets. A moderate amount of flow is accelerated to a high velocity
in these engines. If the exit velocity becomes very high, there are other physical processes which
become important and affect the efficiency of the engine. There is a simplified version of the general
thrust equation that can be used for gas turbine engines. The nozzle of a turbine engine is usually
designed to make the exit pressure equal to free stream. In that case, the pressure-area term in the

53 University of Cambridge, Compressor and Turbine stages.


72

general equation is equal to zero. The thrust is then equal to the exit mass flow rate times the exit
velocity minus the free stream mass flow rate times the free stream velocity54.

4.5 Gas Turbine Performance


The thermodynamic process used in gas turbines is the Brayton cycle. Two significant performance
parameters are the pressure ratio and the firing temperature. The fuel-to-power efficiency of the
engine is optimized by increasing the difference (or ratio) between the compressor discharge
pressure and inlet air pressure. This compression ratio is dependent on the design. Gas turbines for
power generation can be either industrial (heavy frame) or aero derivative designs. Industrial gas
turbines are designed for stationary applications and have lower pressure ratios, typically up to 18:1.
Aero derivative gas turbines are lighter weight compact engines adapted from aircraft jet engine
design which operate at higher compression ratios up to 30:1. They offer higher fuel efficiency and
lower emissions, but are smaller and have higher initial (capital) costs. Aero derivative gas turbines
are more sensitive to the compressor inlet temperature. The temperature at which the turbine
operates (firing temperature) also impacts efficiency, with higher temperatures leading to higher
efficiency. However, turbine inlet temperature is limited by the thermal conditions that can be
tolerated by the turbine blade metal alloy. Gas temperatures at the turbine inlet can be 1200°C to
1400°C, but some manufacturers have boosted inlet temperatures as high as 1600°C by engineering
blade coatings and cooling systems to protect metallurgical components from thermal damage.
Because of the power required to drive the compressor, energy conversion efficiency for a simple
cycle gas turbine power plant is typically about 30 percent, with even the most efficient designs
limited to 40 %. A large amount of heat remains in the exhaust gas, which is around 600˚C as it leaves
the turbine. By recovering that waste heat to produce more useful work in a combined cycle
configuration, gas turbine power plant efficiency can reach 55 to 60 percent. However, there are

Axial
Dynamic
Centrifugal Single Acting

Reciprocating Double Acting


Compressor Types
Diaphram

Positive Displacement Vane

Scroll

Rotery Liquid Ring

Screw

Lobe

Figure 4.9 Gas Compressor Types

54 NASA – Glen Research Center.


73

operational limitations associated with operating gas turbines in combined cycle mode, including
longer startup time, purge requirements to prevent fires or explosions, and ramp rate to full load.

4.6 Gas Compressors


A gas compressor is a mechanical device that increases the pressure of a gas by reducing its volume.
An air compressor is a specific type of gas compressor. Compressors are similar to pumps: both
increase the pressure on a fluid and both can transport the fluid through a pipe. As gases are
compressible, the compressor also reduces the volume of a gas. Liquids are relatively incompressible;
while some can be compressed, the main action of a pump is to pressurize and transport liquids. The
main types of gas compressors are illustrated in Figure 4.9. where here we deal with two commonly
used Axial and Centrifugal compressors.
4.6.1 Axial-Flow Compressors
The dynamic rotating compressors that use
arrays of fan-like airfoils to progressively
compress a fluid. They are used where high
flow rates or a compact design are required.
The arrays of airfoils are set in rows, usually
as pairs: one rotating and one stationary.
The rotating airfoils, also known as blades
or rotors, accelerate the fluid. The
stationary airfoils, also known as stators or
vanes, decelerate and redirect the flow
direction of the fluid, preparing it for the
rotor blades of the next stage (see Figure
4.10). Axial compressors are almost always
multi-staged, with the cross-sectional area
of the gas passage diminishing along the
compressor to maintain an optimum axial
Mach number. Beyond about 5 stages or a Figure 4.10 Schematics of 3.5-stage of Axial
Compressor
4:1 design pressure ratio a compressor will
not function unless fitted with features such
as stationary vanes with variable angles (known as variable inlet guide vanes and variable stators),
the ability to allow some air to escape part-way along the compressor (known as inter-stage bleed)
and being split into more than one rotating assembly (known as twin spools, for example). Axial
compressors can have high efficiencies; around 90% at their design conditions. However, they are
relatively expensive, requiring a large number of components, tight tolerances and high quality
materials. Axial-flow compressors are used in medium to large gas turbine engines, natural gas
pumping stations, and some chemical plants. One can enjoy
the Mach number distribution in the compressor stage of
an axial turbomachine. (see the short animation on the Compressor_Mach_Number-2.mp4
right- Courtesy of Siemens Blog).
4.6.2 Centrifugal Compressors
Centrifugal compressors use a rotating disk or impeller in a shaped housing to force the gas to the
rim of the impeller, increasing the velocity of the gas. A diffuser (divergent duct) section converts the
velocity energy to pressure energy. They are primarily used for continuous, stationary service in
industries such as oil refineries, chemical and petrochemical plants and natural gas processing
plants.[1][14][15] Their application can be from 100 horsepower (75 kW) to thousands of horsepower.
With multiple staging, they can achieve high output pressures greater than 10,000 psi (69 MPa).
Many large snowmaking operations (like ski resorts) use this type of compressor. They are also used
74

in internal combustion engines as superchargers


and turbochargers. Centrifugal compressors are
used in small gas turbine engines or as the final
compression stage of medium-sized gas turbines.
(see Figure 4.11).

4.7 Nomenclature of Terms


Before going further, it is prudent to get
familiarize our self with terminology used in
industry regarding components of
turbomachines . From personal experience, it is
55

an important issue. Some of these are shown in


Figure 4.12 and shown alphabetically in
Figure 4.11 A single stage Centrifugal
Table 4.1 below. Compressor

Table 4.1 Glossary of Turbomachinery Terms

aspect ratio ratio of the blade height to the chord


axial chord Length of the projection of the blade, as set in the turbine, onto a line parallel to the
turbine axis. It is the axial length of the blade.
axial solidity Ratio of the axial chord to the spacing.
adiabatic insulated; occurring with no external heat transfer
blade exit angle Angle between the tangent to the camber line at the trailing edge and the turbine axial
direction.
blade height radius at the tip minus the radius at the hub
blade inlet angle between the tangent to the camber line at the leading edge and the turbine axial
angle direction
blower Rotary machine that produces a low-to-moderate pressure rise in a compressible fluid
(usually air), usually incorporated in a duct. See "fan" and "compressor."
bucket same as rotor blade
camber angle External angle formed by the intersection of the tangents to the camber line at the
leading and trailing edges. It is equal to the sum of the angles formed by the chord line
and the camber-line tangents
camber line Mean line of the blade profile. It extends from the leading edge to the trailing edge,
halfway between the pressure surface and the suction surface
CBE Compressor-burner-expander, or the "simple" gas-turbine "cycle."
CBEX Compressor (heat exchanger)-burner-expander-heat exchanger, or the "regenerated,"
"recuperated," or "heat-exchanger" gas-turbine "cycle."
chord Length of the perpendicular projection of the blade profile onto the chord line. It is
approximately equal to the linear distance between the leading edge and the trailing
edge.

55David Gordon Wilson; "The Design of High-Efficiency Turbomachinery and Gas Turbines", pp 487-492,
published by the MIT Press, Cambridge, Massachusetts, 1984, 5th printing 1991.
75

chord line Two-dimensional blade section were laid convex side up on a flat surface, the chord
line is the line between the points where the front and the rear of the blade section
would touch the surface.
compressor rotary machine that produces a relatively high pressure rise (pressure ratios greater
than 1.1) in a compressible fluid
deflection Total turning angle of the fluid. It is equal to the difference between the flow inlet angle
and the flow exit angle
deviation angle the flow exit angle minus the blade exit angle
diffuser A duct or passage shaped so that a fluid flowing through it will undergo an efficient
reduction in relative velocity and will therefore increase in (static) pressure.
EGV at the exit of the compressor consisting of another set of vanes further diffuses the
fluid and controls its velocity entering the combustors and is often known as the Exit
Guide Vanes (EGV)
effectiveness term applied here to define the heat-transfer efficiency of heat exchangers
efficiency Performance relative to ideal performance. There are many types of efficiency
requiring very precise definitions
entropy A property of a substance defined in terms of other properties. Its change during a
process is of more interest than its absolute value. In an adiabatic process, the increase
of entropy indicates the magnitude of losses occurring
expander A rotary machine that produces shaft power from a flow of compressible fluid at high
pressure discharged at low pressure. Here the only types of expander treated are
turbines
flow exit angle angle between the fluid flow direction at the blade exit and the machine axial direction
flow inlet angle angle between the fluid flow direction at the blade inlet and the machine axial direction
head the height to which a fluid would rise under the action of an incremental pressure in a
gravitational field
hub the portion of a turbomachine bounded by the inner surface of the flow annulus
hub-tip ratio same as hub-to-tip-radius ratio
IGV An additional row of stationary blades that frequently used at the compressor inlet
and are known as Inlet Guide Vanes (IGV) to ensure that air enters the first-stage
rotors at the desired flow angle, these vanes are also pitch variable thus can be
adjusted to the varying flow requirements of the engine
hub-to-tip radio ratio of the hub radius to the tip radius
incident angle the flow inlet angle minus the blade inlet angle
intensive Property that does not increase with mass; for instance, the pressure and temperature
property of a body of material do not double if an equal mass at the same temperature and
pressure is joined to it. (The energy, on the other hand, would double.)
intercoolers heat exchangers that cool a gas after initial compression and before subsequent
compression
isentropic occurring at constant entropy
isothermal occurring at constant temperature
leading edge the front, or nose, of the blade
mean section the blade section halfway between the hub and the tip
meridional a plane cutting a turbomachine through a diametric line and the (longitudinal) axis
plane
nozzle blade same as stator blade, for turbines only
76

pitch the distance in the direction of rotation between corresponding points on adjacent
blades
pressure The concave surface of the blade. Along this surface, pressures are highest
surface
pump A machine that increases the pressure or head of a fluid. In connection with
turbomachinery it usually refers to a rotary machine operating on a liquid.
radius ratio same as hub-to-tip-radius ratio
recuperator a heat exchanger, defined in this book as one with nonmoving surfaces, transferring
heat from a hot fluid to a cold fluid
regenerated See "CBEX."
cycle
regenerator a heat exchanger, defined in this book as one having moving surfaces or valves
switching the hot and cold flows
reheat The effect of losses in increasing the outlet enthalpy, or in decreasing the steam
wetness, in a steam-turbine expansion. Also see "reheat combustor."
reheat a combustor fitted between two turbines to bring the gas temperature at inlet to the
combustor second turbine to approach the temperature at inlet to the first
root The compressor or turbine-blade section attaching it to its mounting platform. Rotor
blade root sections are normally at the hub, and stator- blade roots at the shroud
rotor the rotating part of a machine, usually the disk or drum plus the rotor blades
rotor blade a rotating blade
separation when a fluid flowing along a surface ceases to go parallel to the surface but flows over
a near-stagnant bubble, or an eddy, or over another stream of fluid
shroud the surface defining the outer diameter of a turbomachine flow annulus
solidity the ratio of the chord to the spacing
spacing same as pitch
stagger angle the angle between the chord line and the turbine axial direction (also known as the
setting angle)
stall the condition of operation (usually defined by the incidence) of an airfoil, or row of
airfoils, at which the fluid deflection begins to fall rapidly and/or the fluid losses
increase rapidly
static conditions or properties of fluids as they would be measured by instruments moving
(conditions) with the flow
stator the stationary part of a machine, normally that part defining the flow path
stator blade A stationary blade.
suction surface The convex surface of the blade. Along this surface, pressures are lowest
surge the unstable operation of a high-pressure-ratio compressor whose stalls propagate
upstream from the high-pressure stages or components allowing reverse flow and the
discharge of the reservoir of high-pressure fluid, followed by re-establishment of
forward flow and a repetition of the sequence.
tip The outermost section of the blade or "vane."
total conditions or properties of fluids as they would be measured by stationary
(conditions) instruments that bring the fluid isentropically to rest
trailing edge the rear, or tail, of the blade
transverse the plane normal to the axis of a turbomachine
plane
77

turbine A rotary machine that produces shaft power by extracting energy from a stream of
fluid passing through it, using only fluid-dynamic forces (as distinct from "positive
displacement" or piston-and-cylinder-like machines).
turbomachines As for "turbine," except that the shaft power may be produced or absorbed, and the
energy may be extracted from or added to a stream of fluid.
working fluid Fluid that undergoes compression, expansion, heating, cooling, and other processes in
a heat-engine cycle. In an open-cycle gas turbine the working fluid is air

Figure 4.12 Blade Related Terminology

4.8 Component of Gas Turbine Engine


4.8.1 Inlet
The air inlet duct must provide clean and unrestricted airflow to the engine56. Clean and undisturbed
inlet airflow extends engine life by preventing erosion, corrosion, and Foreign Object Damage (FOD).
Consideration of atmospheric conditions such as dust, salt, industrial pollution, foreign objects (birds,
nuts and bolts), and temperature (icing conditions) must be made when designing the inlet system.
Fairings should be installed between the engine air inlet housing and the inlet duct to ensure
minimum airflow losses to the engine at all airflow conditions. The inlet duct assembly is usually
designed and produced as a separate system rather than as part of the design and production of the

56 “Fundamentals of Gas Turbine Engines”, Cast-Safty.org.


78

engine.
4.8.2 Axial Compressor
The compressor is responsible for providing the turbine with all the air it needs in an efficient
manner. In addition, it must supply this air at high static pressures. The example of a large turboprop
axial flow compressor will be used. The compressor is assumed to contain fourteen stages of rotor
blades and stator vanes. The overall pressure ratio (pressure at the back of the compressor compared
to pressure at the front of the compressor) is approximately 9.5:1. At 100% (>13,000) RPM, the
engine compresses approximately 433 cubic feet of air per second. At standard day air conditions,
this equals approximately 33 pounds of air per second. The compressor also raises the temperature
of the air by about 550F as the
air is compressed and moved
rearward. The power
required to drive a
compressor of this size at
maximum rated power is
approximately 7000
horsepower. In an axial flow
compressor, each stage
incrementally boosts the
pressure from the previous
stage.
A single stage of compression
consists of a set of rotor
blades attached to a rotating
disk, followed by stator
vanes attached to a
stationary ring. The flow area
between the compressor
blades is slightly divergent.
Flow area between
compressor vanes is also
divergent, but more so than
for the blades. In general
terms, the compressor rotor
blades convert mechanical
energy into gaseous energy.
This energy conversion
greatly increases total
pressure (PT). Most of the
increase is in the form of
velocity (V), with a small
increase in static pressure
(PS) due to the divergence of
Figure 4.13 Schematic Diagram of fluid properties through an axial
the blade flow paths. The
compressor stage – Courtesy of [T. B. Ferguson, Gravdahl, and
stator vanes slow the air by Egeland]
means of their divergent duct
shape, converting 'the
accelerated velocity (V) to higher static pressure (PS). The vanes are positioned at an angle such that
the exiting air is directed into the rotor blades of the next stage at the most efficient angle. This
79

process is repeated fourteen times as the air flows from the first stage through the fourteenth stage.
Figure 4.13 shows one stage of the compressor and a graph of the pressure characteristics through
the stage. (see [Niazi]57).
The stator removes swirl from the flow, but it is not a moving blade row and thus cannot add any net
energy to the flow. Rather, the stator rather converts the kinetic energy associated with swirl to
internal energy (raising the static pressure of the flow). Thus typical velocity and pressure profiles
through a multistage axial compressor look like those shown in Figure 4.13. Alternatively, assuming
incompressible, constant density, and with no body force, we can use Bernoulli’s equations (PT = PS +
(1/2)ρV2) where PT is the stagnation pressure, a measure of the total energy carried in the flow, p is
the static pressure a measure of the internal energy, and the velocity terms are a measure of the
kinetic energy associated with each component of velocity58. The rotor adds swirl to the flow, thus
increasing the total energy carried in the flow by increasing the angular momentum (adding to the
kinetic energy associated with the tangential or swirl velocity, 1/2rv2). The stator removes swirl
from the flow, but it is not a moving blade row and thus cannot add any net energy to the flow. Rather,
the stator rather converts the kinetic energy associated with swirl to internal energy (raising the
static pressure of the flow).
Thus a typical velocity and pressure profiles through a multistage axial compressor look like those
shown in Figure 4.14. In addition to the fourteen stages of blades and vanes, the compressor also
incorporates the inlet guide vanes and the outlet guide vanes. These vanes, located at the inlet and
the outlet of the compressor, are neither divergent nor convergent. The inlet guide vanes direct air
to the first stage compressor blades at the "best" angle. The outlet guide vanes "straighten" the air to
provide the combustor with the proper airflow direction. The efficiency of a compressor is primarily

Figure 4.14 Pressure and Velocity profile through a Multi-Stage Axial Compressor

57 Saeid Niazi, “Numerical Simulation of Rotating Stall and Surge Alleviation in Axial Compressors”, A Thesis
Presented to the Academic Faculty, Georgia Institute of Technology, 2000.
58 MIT OpenCourseWare
80

determined by the smoothness of the airflow. During design, every effort is made to keep the air
flowing smoothly through the compressor to minimize airflow losses due to friction and turbulence.
This task is a difficult one, since the air is forced to flow into ever-higher pressure zones. Air has the
natural tendency to flow toward low-pressure zones. If air were allowed to flow "backward" into the
lower pressure zones, the efficiency of the compressor would decrease tremendously as the energy
used to increase the pressure of the air was wasted. To prevent this from occurring, seals are
incorporated at the base of each row of vanes to prevent air leakage. In addition, the tip clearances
of the rotating blades are also kept at a minimum by the use of coating on the inner surface of the
compressor case. All components used in the flow path of the compressor are shaped in the form of
airfoils to maintain the smoothest airflow possible. Just as is the case for the wings of an airplane, the
angle at which the air flows across the airfoils is critical to performance. The blades and vanes of the
compressor are positioned at the optimum angles to achieve the most efficient airflow at the
compressor’s maximum rated speed. Any deviation from the maximum rated speed changes the
characteristics of the airflow within the compressor. The blades and vanes are no longer positioned
at their optimum angles. Many engines use bleed valves to unload the force of excess air in the
compressor when it operates at less than optimum speed.59 The example engine incorporates four
bleed valves at each of the fifth and tenth compressor stages. They are open until 13,000 RPM (~94%
maximum) is reached, and allow some of the compressed air to flow out to the atmosphere. This
results in higher air velocities over the blade and vane airfoils, improving the airfoil angles. The
potential for airfoil stalling is reduced, and compressor acceleration can be accomplished without
surge.
4.8.3 Diffuser
All turbomachines and many other flow systems incorporate a diffuser (e.g. closed circuit wind
tunnels, the duct between the compressor and burner of a gas turbine engine, the duct at exit from a
gas turbine connected to the jet pipe, the duct following the impeller of a centrifugal compressor,
etc.)60. Air leaves the compressor through exit guide vanes, which convert the radial component of
the air flow out of the compressor to straight-line flow. The air then enters the diffuser section of the
engine, which is a very divergent duct. The primary function of the diffuser structure is aerodynamic.
The divergent duct shape converts most of the air’s velocity (Pi) into static pressure (PS) with the aid
of Bernoulli equation. As a result, the highest static pressure and lowest velocity in the entire engine
is at the point of diffuser discharge and combustor inlet. Other aerodynamic design considerations
that are important in the diffuser section arise from the need for a short flow path, uniform flow
distribution, and low drag loss. In addition to critical aerodynamic functions, the diffuser also
provides:

• Engine structural support, including engine mounting to the nacelle


• Support for the rear compressor bearings and seals
• Bleed air ports, which provide pressurized air for:
• Airframe "customer" requirements (air conditioning, etc.)
• engine inlet anti-icing
• control of acceleration bleed air valves
• Pressure and scavenge oil passages for the rear compressor and front turbine bearings.
• Mounting for the fuel nozzles.

59MIT, OpenCourseWare.
60 S. L.
Dixon, “Fluid Mechanics and Thermodynamics of Turbomachinery”, 5th edition, Senior Fellow at University
of Liverpool, 1978-1998.
81

The primary fluid mechanical problem of the diffusion process is caused by the tendency of the
boundary layers to separate from the diffuser walls if the rate of diffusion is too rapid61. The result of
too rapid diffusion is always large losses in stagnation pressure. On the other hand, if the rate of
diffusion is too low, the fluid is exposed to an excessive length of wall and fluid friction losses become
Pre-dominant. Clearly, there must be an optimum rate of diffusion between these two extremes for
which the losses are minimized.
4.8.4 Nozzle
In a large number of turbomachinery components the flow process can be regarded as a purely nozzle
flow in which the fluid receives an acceleration as a result of a drop in pressure (see
Figure 4.7). Such a nozzle flow occurs at entry to all turbomachines and in the stationary blade rows
in turbines. In axial machines the expansion at entry is assisted by a row of stationary blades (called
guide vanes in compressors and nozzles in turbines) which direct the fluid on to the rotor with a
large swirl angle. Centrifugal compressors and pumps, on the other hand, often have no such
provision for flow guidance but there is still a velocity increase obtained from a contraction in entry
flow area. In reality, Nozzle and Diffuser work against each other. A nozzle increases the velocity of a
fluid, while a diffuser decreases the velocity of a fluid. Nozzles can be used by jets and rockets to provide
extra thrust. Conversely, many jet engines use diffusers to slow air coming into the engine for a more
uniform flow.
4.8.5 Combustor
Once the air flows through the diffuser, it enters the combustion section, also called the combustor.
The combustion section has the difficult task of controlling the burning of large amounts of fuel and
air. It must release the heat in a manner that the air is expanded and accelerated to give a smooth and
stable stream of uniformly heated gas at all starting and operating conditions. This task combustion
liners must position and control the fire to prevent flame contact with any metal parts. The engine
under consideration here uses a can-annular combustion section with six combustion liners (cans).
They are positioned within an annulus created by inner and outer combustion cases. Combustion
takes place in the forward end or primary zone of the cans. Primary air (amounting to about one
fourth of the total engine’s total airflow) is used to support the combustion process. The remaining
air, referred to as secondary or dilution air, is admitted into the liners in a controlled manner (Figure
4.15). The secondary air controls the flame pattern, cools the liner walls, dilutes the temperature of

Figure 4.15 Combustor Primary Operating Components

61 See 13.
82

the core gasses, and provides mass. This cooling air is critical, as the flame temperature is above
1930C (3500F), which is higher than the metals in the engine can endure. It is important that the fuel
nozzles and combustion liners control the burning and mixing of fuel and air under all conditions to
avoid excess temperatures reaching the turbine or combustion cases. Maximum combustion section
outlet temperature (turbine inlet temperature) in this engine is about 1070C (>1950F). The rear third
of the combustion liners is the transition section. The transition section has a very convergent duct
shape, which begins accelerating the gas stream and reducing the static pressure in preparation for
entrance to the turbine section.
4.8.6 Axial Gas Turbine
This example engine has a four-stage turbine. The turbine converts the gaseous energy of the
air/burned fuel mixture out of the combustor into mechanical energy to drive the compressor, driven
accessories, and, through a reduction gear, the propeller. The turbine converts gaseous energy into
mechanical energy by expanding the hot, high-pressure gases to a lower temperature and pressure.
Each stage of the turbine consists of a row of stationary vanes followed by a row of rotating blades.
This is the reverse of the order in the compressor. In the compressor, energy is added to the gas by
the rotor blades, then converted to
static pressure by the stator vanes. In
the turbine, the stator vanes increase
gas velocity, and then the rotor blades
extract energy. The vanes and blades
are airfoils that provide for a smooth
flow of the gases. As the airstream
enters the turbine section from the
combustion section, it is accelerated
through the first stage stator vanes.
The stator vanes (also called nozzles)
form convergent ducts that convert
the gaseous heat and pressure energy Figure 4.17 Turbine Flow Characteristics
into higher velocity gas flow (V). In
addition to accelerating the gas, the vanes "turn" the flow to direct it into the rotor blades at the
optimum angle. As the mass of the high velocity gas flows across the turbine blades, the gaseous

Figure 4.16 Schematics of Axial Flow Turbine


83

energy is converted to mechanical energy. Velocity, temperature, and pressure of the gas are
sacrificed in order to rotate the turbine to generate shaft power. Figure 4.17 represents one stage
of the turbine and the characteristics of the gases as it flows through the stage. A multi-stage turbine
is illustrates in Figure 4.15. The efficiency of the turbine is determined by how well it extracts
mechanical energy from the hot, high-velocity gasses. Since air flows from a high-pressure zone to a
low pressure zone, this task is accomplished fairly easily. The use of properly positioned airfoils
allows a smooth flow and expansion of gases through the blades and vanes of the turbine. All the air
must flow across the airfoils to achieve maximum efficiency in the turbine. In order to ensure this,
seals are used at the base of the vanes to minimize gas flow around the vanes instead of through the
intended gas path. In addition, the first three stages of the turbine blades have tip shrouds to
minimize gas flow around the blade tips. We can apply the same analysis techniques to a

Typical Compressor Blade ( Air


Defence Museum) Typical Turbine Blade

Figure 4.18 Examples of Typical Blades for Compressor and Turbine

turbine. Again, the stator does no work. It adds swirl to the flow, converting internal energy into
kinetic energy. The turbine rotor then extracts work from the flow by removing the kinetic
associated with the swirl velocity.

4.9 Difference in
Blading Between Compressor
Compressor and
Turbine • Area increase: pressure rise
There is quite a difference • Flow deceleration: thick boundary layers
between Compressor and • Little flow turning: many stages
Turbine blading. Aside
from shape of it, they are
number of stages and
arrangement of it. While Turbine
Compressor blades are • Area decrease: pressure drop
generally thin and straight,
and resemble a tiny • Flow acceleration: thin boundary layers
rectangular wing with low • Large flow turning: few stages
camber thickness. Turbine
84

blades are more curved. In particularly large and recent engines, where efficiency is critical, turbine
blades will often be full of tiny holes for cooling effects. The difference best described below and
examples of blade shown in Figure 4.18. To distinguish between high pressure and low pressure
stages (compressor or turbine does not matter), the length of the blade and its torsion (i.e. how much
the aerodynamic profile turns around the axis of the blade going from the root to the tip) are key:
shorter and more twisted blades will be high pressure ones, longer and straighter blades will be low
pressure. Note that two blades of the same length could come one from a high pressure stage and the
other from a low pressure one of a different engine: "short" and "long" are relative to the engine size.

4.10 Velocity Triangles in Turbomachines


An important aspect of Turbomachinery is the velocity triangle and their goal to change the flow
apparatus. It is basic vector relationship between relative and absolute frame. Velocity triangles are
typically used to relate the flow properties and blade design parameters in the relative frame
(rotating with the moving blades),
to the properties in the stationary
or absolute frame. It uses the
study of first year Static, and by
“unwrapping” the
compressor. That is, we take a
cutting plane at a particular radius
(see Figure 4.19). Here we have
assumed that the area of the
annulus through which the flow
passes is nearly constant and the
density changes are small so that
the axial velocity is approximately
constant. Let’s examine the
velocities of the gas, as it passes
through a rotor and a stator. At the
point we’re examining, the rotor is
moving with a velocity U. The
velocity of the gas relative to the
rotor is denoted by C and V is Figure 4.19 Velocity triangles for an Axial Compressor
absolute velocity or V = C + U. The
angle between the flow velocity C and the shaft axis is denoted by α. The angle between the rotor
blade angle and the shaft axis is denoted by β. The component of the velocity C in axial direction is
denoted by Ca. It is assumed to be constant along the compressor. Notice the tangential velocity
increase across the rotor for compressor. In some circles, they used W instead of C or W = V – U. In
drawing these velocity diagrams it is important to note that the flow typically leaves the trailing edges
of the blades at approximately the trailing edge angle in the coordinate frame attached to the blade
(i.e. relative frame for the rotor, absolute frame for the stator). We will mainly look at axial
compressors as they are the most used type of compressors. Also, axial compressors work very
similar to axial turbines where stator gives tangential velocity, and rotor moves in the direction of
tangential velocity, having work done on them by flow. Notice tangential velocity decrease across
turbine rotors. (Figure 4.19).

4.11 Energy Exchange with Moving Blades


The Euler turbine equation relates the power added to or removed from the flow, to characteristics
of a rotating blade row. The equation is based on the concepts of conservation of angular momentum
and conservation of energy. They are both turbomachinery: machines that transfer energy from a
85

rotor to a fluid, or the other way around. The working principle of the compressor and the turbine is
therefore quite similar.
4.11.1 Euler’s Equation for Turbomachinery 62
Let’s examine a rotor, rotating at a constant angular velocity ω. The initial radius of the rotor is r 1,
while the final radius is r2. A gas passes through the rotor with a constant velocity c. The rotor causes
a moment M on the gas. The power needed by the rotor is thus P = Mω. It would be nice if we can find
an expression for this moment M. For that, we first look at the force F acting on the gas. It is given by

d(mc)
dFu = = ṁc
dt
Eq. 4.2
Where we have used the assumption that c stays constant. Only the tangential component Fu
contributes to the moment. Every bit of gas contributes to this tangential force. It does this according

dF𝑢 = ṁdc𝑢
Eq. 4.3
Where cu is the tangential velocity of the air. Let’s integrate over the entire rotor. We then find that

2 2 2

M = ∫ dM = ∫ rdFu = ṁ ∫ rdcu = ṁ(cu,2 r2 − cu,1 r1 )


1 1 1
Eq. 4.4
The power is now given by

P = Mω = Tω = ṁ(cu,2 r2 − cu,1 r1 )ω = ṁ(cu,2 r2 − cu,1 r1 )


Eq. 4.5
In this equation, T denotes Torque, u denotes the speed of the rotor at a certain radius r. We have
also used the fact that ω = u1/r1 = u2/r2. The above equation is known as Euler’s equation for
turbomachinery. From Eq. 4.5 it is obvious that:
• If the tangential velocity increases across a blade row (where positive tangential velocity is
defined in the same direction as the rotor motion) then work is added to the flow (a
compressor).
• If the tangential velocity decreases across a blade row (where positive tangential velocity is
defined in the same direction as the rotor motion) then work is removed from the flow (a
turbine).

Furthermore, another form of Euler’s Turbomachinery equation, with aid of the steady flow energy
equation:

H2 − H1 = ω (cu,2 r2 − cu,1 r1 ) = Cp (T2 − T1 ) where Cp = constant Eq. 4.6

It relates the temperature ratio (and hence the pressure ratio) across a turbine or compressor to the
rotational speed and the change in momentum per unit mass. Note that the velocities used in this

62 “Compressor and turbines”, Aero students.


86

equation are what we call absolute frame velocities (as opposed to relative frame velocities).63 It is
given fact that:

• If angular momentum increases across a blade row, then T2 > T1 and work was done on the
fluid (a compressor).
• If angular momentum decreases across a blade row, then T2 < T1 and work was done by the
fluid (a turbine)

Figure 4.20 Velocity Triangles in Relation to Incident Angle

4.12 Compressors and their Reaction to Intake Distortion


During the design phase of an aircraft and its engine it is important that the compatibility aspects at
the aerodynamic interface between the aircraft intake and the engine are given sufficient
consideration because of the implications failures in this area may have64. On a macroscopic level
and in isolation from other effects, the isentropic relation can be applied. Compressors, as the name
implies, compress air by a repeated sequence of first adding kinetic energy to the flow and then
converting the kinetic energy to pressure by a process of flow deceleration. The elements within a
compressor achieving this process are a number of airfoils, either rotating or stationary. Work input
to the flow by a rotor row is achieved via change of the angular momentum of the flow, and these
properties are related to each other via the following equation,

H2 − H1 = u2 c𝑢,2 − u1 c𝑢,1 ⃗⃗⃗⃗ + ⃗U⃗


where c⃗ = w
Eq. 4.7
Especially for axial compressors where rotor angular velocities at rotor inlet and exit are very similar
to each other, it is evident that an increase in total enthalpy requires changing the angular velocity of
flow. Velocity triangles at rotor inlet and exit exemplified in Figure 4.20 show how angular flow
velocities, rotor inlet flow angles and rotor exit flow angles are related to each other. The symbol “c”
denotes velocity in the absolute frame of reference. Due to the rotational speed “u” of the rotor, rotor
blades experience flow velocities within their rotating (or relative) frame of reference, denoted by
the symbol “w”. For the sake of simplicity, it is assumed that the flow at rotor inlet has no angular

63MIT, OpenCourseWare.
64 Breuer, B., Bissinger, N., C., “Encyclopedia of Aerospace Engineering – Volume 8 - Chapter EAE 573-Basic
Principles – Gas Turbine Compatibility – Gas Turbine Aspects”.
87

component (cu, 1 = 0), and the exit flow angle of the rotor blade remains unchanged (in the rotor frame
of reference). With these assumptions, an increase in work input according to equation 1 can only be
achieved by an increase of cu,2. According to the dependencies shown in Figure 4.20, this requires
reducing the axial velocity component of the flow behind the rotor. Because of conservation of mass
flow through the rotor, also the axial velocity at rotor inlet will be reduced, leading to an increased
incidence of the flow to the rotor blade. Translating the state of flow behind the rotor from the
rotating frame of reference into the stationary one, Figure 4.20 also shows that an increase of work
delivered to the flow by the rotor increases the incidence to the subsequent stator row as well.
Therefore, an increase of work input to the flow means increasing incidences to both rotor and stator
airfoils. Therefore, an increase of work input to the flow means increasing incidences to both rotor
and stator airfoils.
Very much like aircraft wings, these airfoils have certain operating limits in terms of airfoil angle of
attack or incidence. With increasing incidence, rotor airfoils provide for a larger work input and
hence pressure rise, but at the same time the aerodynamic loading increases, up to a point where the
flow separates. On a larger scale, the pressure rise capability of a compressor is typically depicted
using a compressor map where pressure rise is depicted as a function of compressor mass flow for
different rotational speeds.
An example map is provided
with Figure 4.21, and for the
sake of illustration, it also
relates different regimes of
compressor operating range
to an aircraft operating at
different angles of attack. At
low pressure ratios, the
airfoils operate with negative
to small incidence, and usually
elevated losses. When the
pressure ratio is increased,
airfoil incidences now
approach a condition with
minimum losses. Further
increasing the pressure ratio
is equivalent to further rise of
airfoil aerodynamic loading
and losses increase due to Figure 4.21 Compressor Operating Map
formation of regions of
separated flow. At the upper end of a speed line, there is a point where regions of separated flow have
enlarged to an extent where no further pressure rise is achievable, in analogy to aircraft wings
reaching the stall limit where no further increase of lift can be provided65. The upper operational
limit of a compressor map is called the surge line, representing a condition where large flow
separation prevents further pressure rise. The surge line represents an operational limit for engine
operation, since the occurrence of compressor surge (sometimes also referred to as compressor stall)
leads to a highly unsteady flow field within the engine, quite often also entailing periods of reversed
flow, that is air flowing in the “wrong” direction through the compressor. Surge is associated with
large fluctuations of power output. Furthermore, it is accompanied by increased structural loads
caused by the rapid changes of flow field state. Compressor maps are usually established (either
numerically or by means of testing) for a standard set of inlet conditions. These inlet conditions are

65 See 71.
88

typically derived from simplified installation assumptions and assume a simplified inlet profile with
radial variations only, but uniform in circumferential direction. Intake distortion considerations deal
with conditions that deviate from these design assumptions and aim to identify the consequences of
these deviations with regard to engine operation.

4.13 Effects of Turbine Temperature


The materials used in the turbine section of the engine limit the maximum temperature at which a
gas turbine engine can operate66. The first metal the hot gases from the combustion section strike is
the turbine inlet. The temperature of the gas stream is carefully monitored to ensure that over
temperature does not occur. Compromises are made in turbine design to achieve the optimum
balance of power, efficiency, cost, engine life, and other factors. As an example, our sample engine
can operate at a higher turbine inlet temperature than previous models due to improved materials
and design. The higher temperature allows for increased power and improved efficiency while
adding higher cost for the direct cooling of the first turbine stage airfoils and other components.
Figure 4.22 shows the temperature, velocity and pressure variation across a gas turbine engine67.
To increase the overall performance of the engine and reduce the specific fuel consumption, modern
gas turbines operate at very high temperatures. However, the high temperature level of the cycle is
limited by the melting point of the materials. Therefore, turbine blade cooling is necessary to reduce
the blade metal temperature to increasing the thermal capability of the engine. Due to the
contribution and development of turbine cooling systems, the turbine inlet temperature has doubled
over the last 60 years. The cooling flow has a significant effect on the efficiency of the gas turbine. It
has been found that the thermal efficiency of the cooled gas turbine is less than the uncooled gas
turbine for the same input conditions (Figure 4.23). The reason for this is that the temperature at
the inlet of turbine is decreased due to cooling and therefore, work produced by the turbine is slightly
decreased. It is also known that the power consumption of the cool inlet air is of considerable concern

Pressure Temperature Velocity

Figure 4.22 Sample engine Perssure, Velocity and Temperature variation

66“Fundamentals of Gas Turbine Engines”, Cast-Safety.org.


67Shahrokh Sorkhkhah, “Gas Turbine Fundamentals”, Faculty of Karaj , Azad University Design Director of Iran
Gas Turbine Company.
89

since it decreases the net power output of the gas turbine68-69. With this in mind, during the design
phase of gas turbine it is very important to optimize the cooling flow if you are considering both the
performance and reliability. Cooled Gas turbine design is quite complicated and requires not only the
right methodology, but also the most appropriate design tools, powerful enough to predict the results
accurately from thermodynamics cycle to aerothermal design, ultimately generating the 3D blade.
Different cooling methods that are employed depend on the extent of the cooling required. The
cooling flow passes through several loops internally and is then ejected over the blade surface to mix
with the main flow. The mixing of the cooling flow with the main flow alters the aerodynamics of the
flow within the turbine cascade. The cooling flow that is injected into the main flow needs to be
optimized, not only in terms of thermodynamic parameters, but also in terms of the locations to
ensure the turbine vanes and blade surfaces are maintained well below the melting surface. The
spacing between the holes, both in horizontal and vertical direction, affects not only the surface
temperature of the blade, but also the strength of the blade and its overall life.
Performing a 3D analysis for optimizing the flow, spacing, and location of cooling flow is
computationally expensive. One has to resort to reduced order 1D flow and heat network simplifies
the task of not only arriving at the optimal configuration of cooling holes and location, but also in
aerothermal design of the gas turbine flow path and generation of the optimized 3D blades with
reduced overall design cycle time. Designers are faced with the challenge of simplifying the complex
3D cooled blade and accurately modelling it.

Figure 4.23 Turbine Inlet Temperature27

68 Amjed Ahmed Jasim AL-Luhaibi, Mohammad Tariq, “Thermal Analysis of Cooling Effect on Gas Turbine Blade”,
eISSN: 2319-1163 | pISSN: 2321-7308.
69 Posted by: Abdul Nassar, “Optimizing the Cooling Holes in Gas Turbine Blades”, SoftInWay® Incorporated,

2016.
90

4.14 Compressor and Turbine Characteristics


The compressor has several important parameters. There are the mass flow m͘ , the initial and final
temperatures T02 and T03, the initial and final pressures p02 and p03, the shaft speed ω (also denoted
as N), the rotor diameter D, and so on70. Let’s suppose we’ll be working with different kinds of
compressors. In this case, it would be nice if we could compare these parameters in some way. To do
that, dimensionless parameters are used. By using dimensional analysis, we can find that there are
four dimensionless parameter groups. They are

ṁ√RT02 p03 ωD
, , and η
p02 D2 p02 √RT02
Eq. 4.8
These parameter groups are known as the mass flow parameter group, the pressure ratio, the
shaft speed parameter group and the efficiency. The efficiency can be either polytrophic or
isentropic. (These two efficiencies depend on each other anyway). The relation between the four
dimensionless parameters can be captured in a graph, known as a characteristic. An example of a
characteristic is shown in Figure 4.24. When applying dimensional analysis to a turbine, the same
results will be found. However, this time the initial and final pressures are p04 and p05. The initial and
final temperatures are T04 and T05.
4.14.1 Stall
Let’s examine the air entering the
rotor71. Previously, we have
assumed that this air has exactly
the right angle of incidence “i” to
follow the curvature of the rotor
blade. In reality, this is of course
not the case. In fact, if the angle of
incidence is too far off, then the
flow can’t follow the curvature of
the rotor blades. The other
phenomena associated with Stall is
if there are pockets of low axial
velocity covering one or two blade
passages (see Figure 4.26). This
is called stall and usually starts at
one rotor blade. However, this stall
alters the flow properties of the air
around it. Because of this, stall Figure 4.24 Characteristics Graph of a Compressor
spreads around the rotor. And it
does this opposite to the direction of rotation of the rotor. This phenomenon is called rotating stall.
Often, only the tips of the rotor blades are subject to stall. This is because the velocity is highest there.
This is called part span stall. If, however, the stall spreads to the root of the blade, then we have full
span stall. For high compressor speeds ω, stall usually occurs at the last stages. On the other hand,
for low compressor speeds, stall occurs at the first stages. Generally, the possibility of stalling
increases if we get further to the left of the characteristic. (See also Figure 4.24).

70 MIT OpenCourseWare.
71 See previous.
91

4.14.2 Compressor Surge72


Let’s suppose we control the mass flow m˙ in a
compressor, running at a constant speed ω. The
mass flow m˙ effects the pressure ratio p03/p02.
There can either be a positive or a negative
relation between these two. Let’s examine the
case where there is a negative relation between
these two parameters. Now let’s suppose we
increase the mass flow m˙. The pressure at the
start of the compressor will thus decrease.
However, the pressure upstream in the
compressor hasn’t noticed the change yet.
There is thus a higher pressure upstream than
downstream. This can cause flow reversal in the
compressor. Flow reversal itself is already bad.
However, it doesn’t stop here. The flow Figure 4.25 Classical Compressor Surge Cycles
reversal causes the pressure upstream in the
compressor to drop. This causes the compressor to start running again. The pressure upstream again
increases. Also, the mass flow increases. But this again causes the pressure downstream to increase.
Flow reversal thus again occurs. A rather unwanted cycle has thus been initiated. This cyclic
phenomenon is called surge. It causes the whole compressor to start vibrating at a high frequency
(see Figure 4.25). Surge is different from stall, in that it effects the entire compressor. However,
the occurrence of stall can often lead to surge. There are several ways to prevent surge. We can blow-
off bleed air. This happens halfway through the compressor. This provides an escape for the air.
Another option is to use variable stator vanes (VSVs). By adjusting the stator vanes, we try to make
sure that we always have the correct angle of incidence i. Finally, the compressor can also be split up
into parts. Every part will then
have a different speed ω.
Contrary to compressors,
turbines aren’t subject to
surge. Flow simply never tends
to move upstream in a turbine.
As an alternative, an interesting
read regarding stall and surge of
compressors is presented by
[Niazi]73.
4.14.3 Choked Flow
Let’s examine the pressure ratio
p04/p05 in a turbine. Increasing
this pressure ratio usually leads
to an increase in mass flow m˙.
However, after a certain point,
the mass flow will not increase
further. This is called choked Figure 4.26 Illustration of the Propagation of a Stall Cell in the
flow74. It occurs, when the flow Relative Frame

72 MIT OpenCourseWare.
73 Saeid Niazi, “Numerical Simulation of Rotating Stall and Surge Alleviation in Axial Compressors”, A Thesis
Presented to the Academic Faculty of Georgia Institute of Technology, 2000.
74 See previous.
92

reaches supersonic velocities. Choked flow can also occur at the compressor. If we look at the right
side of Figure 4.24, we see vertical lines. So, when we change the pressure ratio p03/p02 at constant
compressor speed ω, then the mass flow remains constant.

4.15 Additional Types of Turbines


Beside gas turbine which was the main concern here, there are other types of turbine used in
industry, namely Wind, Steam and Hydraulic turbines (see Figure 4.1). Their main purpose is to
harness the useful mechanical energy (electricity) from kinetic and potential energies. These can be
characterized as renewable energy category. For types of wind turbine, readers are encourage to
consult the work by [Ragheb]75, or Introduction to wind turbine Aerodynamics, Springer 2014, or
section Error! Reference source not found. of this publication, among others. For Steam and H
ydraulic turbines, an excellent references are given by Wikipedia.

75Magdi Ragheb and Adam M. Ragheb (2011). “Wind Turbines Theory - The Betz Equation and Optimal Rotor
Tip Speed Ratio, Fundamental and Advanced Topics in Wind Power”, Dr. Rupp Carriveau (Ed.), ISBN: 978-953-
307-508-2, InTech, Available from: http://www.intechopen.com/books/fundamental-and-advanced-topicsin-
wind-power/wind-turbines-theory-the-betz-equation-and-optimal-rotor-tip-speed-ratio.
93

5 Primary Research in Turbomachinery


5.1 Research Spectrum
The design of turbomachinery is a complex task due to the complicated flow phenomena and
interaction of multi-disciplines which involves aerodynamics, heat transfer76, structural dynamic,
control theory, materials and manufacture engineering etc. Among these design processes,
aerodynamic analysis is the keystone of the design, which decides the performance of
turbomachinery directly.
While, without numerical
technologies (CFD simulation
and numerical optimization), it
is impossible to meet the
increasing rigorous
requirements of design. Hence,
the research on numerical
aerodynamic analysis and
numerical design of
turbomachinery are
outstandingly important. The
aerodynamic performance of
turbomachinery mainly
depends on the complex
internal flows which usually
are strongly three dimensional,
viscous and unsteady. Figure
5.1 shows the impact of CFD on
SNECMA fan performance over
a 30 year period. The flows in
blade passages may be laminar,
turbulent and transitional, and Figure 5.1 Impact of CFD on SNECMA fan performance, over a
may include wake flow, and period of 30 years
secondary flows etc. In
addition, there also may exist other complicated flow phenomena, such as transition, boundary layer
separation, shock and shock-boundary layer interaction, the unsteady interaction between the blade
rows, the interactions between the blade row and end-wall, etc. In 1999, a NASA report of “Numerical
Simulation of Complex Turbomachinery Flows”77 stated four typical complex flows in
turbomachinery which have been investigated extensively and may remain being the key research
problems of turbomachinery in next few decades. These flows are:
• Unsteady flow
• Turbulence
• Film cooling
• Three dimensional flow in turbine including tip leakage effect

76 Xiaodong Wang, “CFD Simulation of Complex Flows in Turbomachinery and Robust Optimization of Blade
Design”, Submitted to the Department of Mechanical Engineering Doctor of Philosophy at the Vrije Universiteit
Brussel July 2010
77 X. D. Wang, Sh. Kang, “Solving stochastic burgers equation using polynomial chaos decomposition”, J. Eng.

Therm., 31(3):393-398, 2010. (In Chinese)


94

5.2 Application of CFD in Turbomachinery


Accurate and robust turbomachinery off-design performance prediction remains elusive.
Representation of transonic compression systems, most notably fans, is especially difficult, due in
large part to highly three-dimensional blade design and the resulting flow field78. Complex shock
structure and subsequent interactions (with blade boundary layers, end-walls, etc.) provide
additional complications. Surely, turbomachinery design has benefited greatly from advancements
in computational power and efficiency. However, practical limitations in terms of computational
requirements, as well as limitations of turbulence and transition modeling, make it difficult to use
CFD to analyze complex off-design issues. Accurate and robust turbomachinery off-design
performance prediction remains elusive. Representation of transonic compression systems, most
notably fans, is especially difficult, due in large part to highly three-dimensional blade design and the
resulting flow field. Complex shock structure and subsequent interactions (with blade boundary
layers, end-walls, etc.) provide additional complications. Surely, turbomachinery design has
benefited greatly from advancements in computational power and efficiency. However, practical
limitations in terms of computational requirements, as well as limitations of turbulence and
transition modeling, make it difficult to use CFD to analyze complex off-design issues.
For example, CFD analyses have only recently been used to explore the complex flow fields resulting
from inlet distortion through modern multistage fans. The time-accurate investigation by Hah, et al,
1998, which included unsteady circumferential and radial variations of inlet total pressure, is one of
the most complete in the open literature. Even so, Hah’s calculation was limited to two blade passages
with boundary conditions just upstream and downstream of the first rotor of a two-stage fan. As
discussed below, improvements to traditional numerical approaches are needed. With the
development of computer technology, the Reynolds Averaged Navier-Stokes (RANS) simulations are
developed rapidly since 1980s. In the same time, a couple of turbulence models are proposed
successively to complete RANS model. In most design processes, the steady RANS simulations give
satisfied prediction of overall performance. While in elaborate design processes, unsteady RANS
(URANS) simulations are needed since the flows in turbomachinery are highly unsteady. Respecting
to the approximation level of geometry, CFD simulation of turbomachinery developed from 2D to 3-
D, from planar cascade to annular cascade, from single blade passage to whole ring, from single stage
to multi stages. The increase of model accuracy to the real geometry has significant effects on
turbomachinery design. Figure 5.1 Exhibits the impact of CFD on the performance improvement of
aircraft engine in SNECMA (France) over a period of almost 30 years79. The evolution, from the initial
use of simple 2D potential flow models in the early 1970s to the current applications of full 3D Navier-
Stokes code, has led to overall gain in efficiency close to 10 points80.

5.3 Quasi 3D Flow (Q3D)


The definition Fully 3D methods replace the stream surface calculation of blade-to-blade (S1) and
hub-to-tip (S2) stream surface was introduced by Wu81 , and this viewpoint dominated the subject
until the early 1980s when fully three dimensional (3D) methods first became available. Wu’s static
pressure S1/S2 approach was far ahead of his time in that he saw flow velocity it as a method of

78 Boyer, K., M., “An Improved Streamline Curvature Approach for Off-Design Analysis of Transonic Compression
Systems “, PhD. Dissertation, Virginia Polytechnic Institute and State University, 2001.
79 J. F. Escuret, D. Nicoud, and Ph. Veysseyre,”Recent advances in compressors aerodynamic design and analysis”,

AVT TP/1, RTO/NATO, 1998.


80 Xiaodong Wang, “CFD Simulation of Complex Flows in Turbomachinery and Robust Optimization of Blade

Design”, Submitted to the Department of Mechanical Engineering Doctor of Philosophy at the Vrije universiteit
Brussel July 2010.
81 Wu, C. H. “A general through flow theory of fluid flow”, NACA paper TN2302, 1951.
95

solution for fully 3D flow. Wu’s ideas


were considerably simplified by
circumferential distance assuming that
the S1 stream surfaces were surfaces of
revolution (i.e. untwisted) while the S2
stream surfaces were reduced to a single
mean stream surface that could be
treated as an axisymmetric flow (Figure
5.2). The axisymmetric hub-to-tip (S2)
calculation is often called the ‘Through
flow calculation’ and has become the
backbone of turbomachinery design,
while the ‘blade-to-blade’ (S1) calculation
remains the basis for defining the
detailed blade shape. Fully 3D methods
replace the stream surface calculation of
blade-to-blade (S1) and hub-to-tip (S2)
stream surface calations by a single Figure 5.2 Illustration of S1 and S2 surfaces
calculation for the whole blade row. This
removes the modelling assumptions of the quasi 3D (Q3D) approach but requires far greater
computer power and so was not usable as a design tool until the late 1980s. For similar reasons, early
methods had to use coarser grids that introduced larger numerical errors than in the Q3D approach.
Radial equilibrium and through-flow methods determine the meridional variations in the velocity
field, but they assume that the turbomachinery flow field is axisymmetric. Cascade analysis and
blade-to-blade computational methods consider the flow variations across the blade passages, but
they neglect span wise variations and radial flows. These two views of a turbomachine are very useful
and both are essential in the design process, but in reality the flow field in all axial turbomachinery,
to some degree, varies in the axial, radial, and tangential directions.
5.3.1 Stream Surface of Second Kind - Through Flow (S2)
Through flow calculations can be used in design (or inverse) mode to determine blade inlet and exit
angles and velocity variation from a specified span-wise work distribution, or in analysis (or direct)
mode when blade angles are specified and flow angles, work, and velocity distributions are predicted.
Through flow calculation programs are probably the most important tool of the turbine aerodynamic
designer. At the initial design stage a one-dimensional mean line calculation might be used to obtain
estimates of blade height and so to lay out a first approximation to the annulus line. Such mean line
calculations usually include estimates of blade loss and deviation, so that predictions of turbine
performance can be obtained, but these must be based only on the blade geometry mid-height so
high accuracy cannot be expected. Although span wise variations in flow are small for very high
radius ratio turbines these variations become significant at radius ratios below about 0.9. It is well
known that most turbine blades are remarkably tolerant to off-design incidences (compared to
compressor blades), but even so optimum performance, particularly at off-design conditions, cannot
be expected unless the blades are matched to the span wise variation in flow. The main objective of
a through flow calculation is, therefore, to provide a prediction of this span wise variation so that
suitable blade profiles can be selected to cope with the variations in inlet angle, turning, Mach
number, etc. The main problem encountered when developing through flow calculations for turbines,
as opposed to compressors, arises from the need to be able to calculate the flow through stages with
high-pressure ratio and in particular with regions of transonic relative flow. The latter is much more
easily handled by Streamline Curvature methods (SCM) than by stream function methods although
severe difficulties arise even for the former type of method. Time-marching methods are much better
96

suited to calculating transonic flow but are not yet highly developed further use in through flow
calculations. Problems with calculating transonic flow are currently much more severe in steam
turbines than in gas turbines. The traditional use of streamline curvature method (SCM)
approaches, as most often discussed in the literature during the preliminary design phase, are
discussed in detail in82. The stream surface represented by

s(r, ψ, z) = 0
Eq. 5.1
As depicted in Figure 5.3. The through-flow solver provides a preliminary blade shape, continually
refined through solutions from higher-order and secondary flow models. One way to calculate a 3D
flow field is to solve two sets of equations, one dealing with axis-symmetric flow in the meridional
plane, commonly referred to as the “S2” surface, and the other with blade-to-blade flow on a stream
surface of revolution, the “S1” plane (see Figure 5.2). The traditional formulation for the governing
momentum equation(s) is a first-order velocity gradient representation, one in the radial and one in
the tangential direction approach for off-design analysis along an axis-symmetric S2 surface. It is
generally accepted that any streamline curvature solution technique will yield satisfactory flow
solutions as long as the deviation, losses, and blockages are accurately predicted83.

Figure 5.3 Streamline Curvature Method

5.3.2 Stream Surface of First Kind (Blade 2 Blade – S1)


These methods calculate the flow on a blade-to-blade (S1) stream surface given the stream surface
shape with the objective with an associated stream surface thickness and of designing the detailed
blade profile. The stream surface is best thought of as a stream tube radius which are obtained from
the through flow calculation. Accurate specification of the radius and thickness variation is essential
as they can have a dominant effect on the blade surface pressure distribution. As with through flow
methods the calculation may be in either direct (or analysis) mode, when the blade shape is
prescribed and its surface pressure distribution calculated, or in inverse mode, where the required
blade surface pressure distribution is prescribed and a blade shape is sought. Many different
numerical methods have been developed for this task. Initially streamline curvature (to be
discussed later) and stream function methods were popular, but both have difficulty coping be made

82 Chung-Hua Wu, “A General Theory of Three Dimensional Flow in Subsonic and Supersonic Turbomachines of
Axial-Radial- and Mixed Flow Types”, National Advisory Committee for Aeronautics, Technical Note, 1952.
83 Boyer, K., M., “An Improved Streamline Curvature Approach for Off-Design Analysis of Transonic Compression

Systems “, PhD. Dissertation, Virginia Polytechnic Institute and State University, 2001.
97

to calculate transonic flows with weak shock with transonic flow and they have now largely been
abandoned. Velocity potential methods can waves but they have seen limited use in turbomachinery.
The numerical methods described above are inviscid and need to be coupled to a boundary layer
calculation if they are to be used to predict blade loss. For com pressor blades the boundary layer
blockage must be included in the inviscid calculation as it significantly affects the blade surface
pressure distribution84. For most turbine blades the boundary layer is so thin that it may be
calculated separately after obtaining the surface pressure distribution from an inviscid calculation. A
recent alternative (N–S) equations which predict the boundary native to coupled inviscid/boundary
layer calculations is the direct solution of the Navier– layer growth as part of the main calculation.
These demand a much finer grid near to the blade surfaces than do inviscid calculations and so are
considerably more ‘expensive’. Nevertheless the N–S equations for blade to-blade flow are now
routinely solved as part of the design process, requiring only a few minutes CPU time on a modern
workstation. There remains controversy about the best turbulence and transition models to use and
about how many mesh points are necessary within the boundary layer.
A variety of blade–to-blade solvers are currently available in the design system. They range from
potential and streamline curvature method up to fully viscous, time marching solvers. The main use
of the blade-to-blade codes is to ensure that the vector diagrams set by thorough Flow are achievable
and within the bounds of blade thickness, loading and efficiencies. For examples, in turbine design
the suction surface diffusion is taken as a primary indicator as to the condition of the boundary layer.
The blade-to-blade code solves for the suction surface velocity ratio, or diffusion factor, and the
geometry is adjusted accordingly. Most of these codes are very similar to those available in other
design systems and have also been described elsewhere. However, three codes (TAYLOR, AEGIS and
NOVAKED2D) are different and worth mentioning85.
5.3.3 Case Study – Turbine Airfoil Optimization Using Inviscid Quasi 3D (Q3D) Analysis Codes
Citation : Ng, E., & Yi, M. (1998). Computation of Q3D Viscous Flows in Various Annular Turbine Stages
with Heat Transfer. International Journal of Rotating Machinery, 4, 25-33.
Turbine airfoil design has long been a domain of expert designers who use their knowledge and
experience along with analysis codes to make design decisions86. The turbine aerodynamic design is
a three-step process that is pitch line analysis, through-flow analysis, and blade-to-blade analysis, as
depicted in Figure 5.4. In the pitch line analysis, flow equations are solved at the blade pitch, and a
free vortex assumption is used to get flow parameters at the hub and the tip. Using this analysis the
flow path of the turbine is optimized, and number of stages, work distribution across stages, stage
reaction, and number of airfoils in each blade row are determined. In the through-flow analysis, the
calculation is carried out on a series of meridional planes where the flow is assumed to be
axisymmetric and the boundary conditions of each stage are determined. The axisymmetric through-
flow method allows for variation in flow parameters in the radial direction without using the free
vortex assumption and accounts for interactions between multiple stages. In the blade-to-blade
analysis, airfoil profiles are designed on quasi-3D surfaces using a computational fluid dynamics
code.
The primary sources of losses in an airfoil are profile loss, shock loss, secondary flow loss, tip
clearance loss, and end-wall loss. Profile loss is associated with boundary layer growth over the blade
profile causing viscous and turbulent dissipation. This also includes loss due to boundary layer

84 Calvert, W. J. and Ginder, R. B., “Quasi-3D calculation system for the flow within transonic compressor blade
rows”, ASME paper 85-GT-22, 1985.
85 Ian K. Jennions, “Elements of a Modern Turbomachinery Design System”, GE Aircraft Engines, One Neumann

Way, MD X409, Cincinnati, OH 45215-6301,United States.


86 E.Y.K. NG, and MIAO YI , “Computation of Q3D Viscous Flows In Various Annular Turbine Stages With Heat

Transfer”, International Journal Of Rotating Machinery, 1998, Vol. 4, No. 1, pp. 25-33.
98

separation because of conditions such as extreme angles of incidence and high inlet Mach number.
Shock losses arise due to viscous dissipation within the shock wave which results in increase in static
pressure and subsequent thickening of the boundary layer, which may lead to flow separation
downstream of the shock. End-wall loss is associated with boundary layer growth on the inner and
outer walls on the annulus. Secondary flow losses arise from flows, which are present when a wall
boundary layer is turned through an angle by an adjacent curved surface. Tip clearance loss is caused
by leakage flows in the tip clearance region of the rotor blade, where the leaked flow fails to
contribute to the work output and also interacts with the end-wall boundary layer. The objective of
the design is to create the most efficient airfoil by minimizing these losses. This often requires
trading-off one loss versus another such that the overall loss is minimized.
To compute all these losses a 3D viscous analysis is required; however, due to the computational load
of such a code, a quasi-3D analysis code is often used in the design process. Thus the impact of the
blade geometry on 3D losses cannot be determined and only 2D losses can be minimized, that is,
profile and shock losses. A viscous quasi-3D analysis though less computationally intense is still too
expensive for use in design optimization, and an inviscid quasi-3D code is used instead. Consequently,
viscous losses are not computed from the analysis code and airfoil performance is gauged by the
characteristics of the Mach number distribution on the blade surface. The most practical formulation
for low-speed turbine airfoil designs still remains the direct optimization formulation based on 2D
inviscid blade-to-blade solvers.

Figure 5.4 The turbine Design Process

5.3.3.1 Quasi-3D CFD Analysis and Results


A quasi 3D CFD solver is used in the current investigation to analyze the flow on the airfoil, which is
an isentropic that uses the streamline curvature method that computes the Mach Number/Pressure
distribution on the airfoil surface87. In the absence of a viscous code, designers usually estimate the
quality of the airfoil by visually examining the Mach number distribution obtained from an in-viscid
quasi 3D CFD solution. Since optimization techniques are driven by a numerical value of the objective

87 Goel, Sanjay,” Turbine Airfoil Optimization Using Quasi-3D Analysis Codes”, University at Albany, USA.
99

function, and the visual perspective of the designer is the only proven metric available, it must be
captured in a suitable numerical algorithm to provide a measure of quality of an airfoil. The current
work employs curve fitting coupled with design heuristics to compute quality metrics from the Mach
number distribution and the airfoil geometry.
These metrics are weighted for different designs based on individual designer preferences. Primary
evaluation metrics that have been defined are diffusion, deviation, incidence deviation, and leading
edge crossover. A physical interpretation of these metrics is presented below. Diffusion is defined as
the deceleration of the flow along the blade surface. It is measured as the cumulative aggregate of all
flow diffusions at each point along the airfoil surface. As the flow diffuses, the boundary layer
thickens, and the momentum loss in the boundary layer increases. In this case, the increased drag
causes a significant loss of momentum; flow separation may result, causing much larger losses. Thus,
the objective of the design is to minimize the diffusion effect. Since the impact of diffusion on the
pressure and suction sides is different, separate terms are defined for the suction and pressure sides.
In the test case presented here, a low-pressure turbine nozzle is optimized. The flow-path of the low-
pressure turbine used in the investigation is shown in Figure 5.5. The radial distances in the figure
are measured with reference to the centerline of the engine and the axial distances are measured
with reference to a point upstream of the first stage of the turbine. The horizontal lines in the figure
represent the streamlines of the flow. Thirteen streamlines are shown, the top and bottom of which
coincide with the casing and the hub respectively. The vertical lines represent the edges of the blade
rows and the location of the frame. The turbine has six stages, each stage composed of two blade
rows. The first blade row consists of nozzles and the second blade row consists of buckets. The stages
are numbered from 1 to 6 in the Figure 5.5.

Figure 5.5 Flow Path of the Turbine


In the current investigation, stage 5 nozzle was designed using sections from five streamlines equally
spaced along the blade span (hub to tip). Figure 5.6 Shows the approximate locations of the
streamlines for an airfoil in which the first and the last streamlines are shown at the hub and tip. In
reality however streamlines at 5% and 95% span were used instead of streamlines directly on the
hub and tip because Mach number distributions very close to the end walls are distorted by the end
wall effects and not representative of the flow away from the walls. The starting solution for the test
case was obtained by estimating the airfoil shape based on shapes of similar airfoils designed in the
past. All the Mach number and airfoil geometry plots use the same reference radial and axial
100

locations as shown in Figure 5.7. To ensure slope and curvature smoothness of the geometry,

Figure 5.7 3D model of an


Figure 5.6 Schematics of an airfoil showing stream airfoil showing the passage
lines along the radial direction between adjacent airfoils

second- order polynomials were used to represent the radial distribution of geometry parameters.

5.4 Theory of Radial Equilibrium in Through Flow (Cr = 0)


Consider a small element of fluid of mass dm shown in Figure 5.8 of unit depth and subtending an
angle dθ at the axis, rotating about the axis with tangential velocity, cθ, at radius r. The element is in
radial equilibrium so that the pressure forces balance the centrifugal forces (cr = 0):

1 dmcθ2
(p + dp) (r + dr) dθ − p r dθ − (p + dp) dr dθ =
2 r
Eq. 5.2
Writing dm = ρ r dϴ dr and ignoring terms of 2nd order we obtain:

1 dp cθ2
=
ρ dr r
Eq. 5.3
For an incompressible fluid and using
thermodynamic relations the Radial Equilibrium
Equation can be written as:

dh 0 ds
− T = cx x + θ
dc c d
(rcθ ) or
dr dr dr r dr
dc
cx x + θ
c d
(rcθ ) = 0
dr r dr
1 dp 0 1 dp dc dc
= + c x x + cθ θ or Figure 5.8 Radial Equilibrium
ρ dr ρ dr dr dr
1 dp 0 dc
= cx x + θ
c d
(rcθ )
ρ dr dr r dr
Eq. 5.4
101

This equation clearly states that equal work is delivered at all radii and the total pressure losses
across a row are uniform with radius. It may be applied to two sorts of problem: the design (or
indirect) problem, in which the tangential velocity distribution is specified and the axial velocity
variation is found, or the direct problem, in which the swirl angle distribution is specified, the axial
and tangential velocities being determined.

5.5 Governing Equation of Rotating Frame of Reference


Accounting for the particular flow situation in turbomachinery, it is necessary to be able to describe
the flow behavior relatively to a rotating frame of reference that is attached to the rotor. Without loss
of generality, it is assumed that the moving part of turbomachinery is rotating steadily with angular
velocity ω around the machine axis along which a coordinate z is aligned. Define u as absolute
velocity, w is relative velocity, and v is as rotating system or blade ω ⨯ r, we have,


𝐮 = 𝐰
⏟ + ⏟
𝐯 =𝐰+𝛚×𝐫
Absolute Relative Coordinates
Eq. 5.5
Introducing this into the mass conversation and after some manipulation we obtain,

∂r ρ
+ ∇. (ρ𝐰) = 0
∂t
Eq. 5.6
Comparing with non-inertia frame of reference, it seems to keep the same expression where
subscript r refers to the rotating frame of reference. Without causing confusion, the subscript r can
be omitted in general. The total derivative (acceleration) is also can be redefined as

Du ∂w ∂v
= + + w. (∇w) + 2 w
⏟× ω + ω ⏟× v
Dt ∂t ∂t Coriolis Centrifugal
Eq. 5.7
The first item on right-hand side expresses the local acceleration of the velocity field within the
rotating frame of reference. The second term and third item denote the angular velocity acceleration

Figure 5.9 Coriolis and Centripetal forces created by the Rotating Frame of Reference
102

and the convective term within the rotating frame of reference, respectively. While, the fourth item
and last item are the Coriolis acceleration and the Centrifugal acceleration, respectively, which are
fictitious forces produced as a result of transformation from stationary frame to rotating frame of
reference. Figure 5.9 shows the directions of the velocity and the acceleration, and relationship
between the absolute velocity, relative velocity and rotation (Schobeiri, 2005). Substituting the
acceleration in Error! Reference source not found. distinctly, for an incompressible flow equations o
f motion and energy, in rotating frame of reference can be obtained:

 ( w )  ( v)
Momentum : + + w.(w ) + ω  v + 2ω  w = μw − p + F
t t
  w
2
v 
2

D ρ h + + 
  2 2   p
=
Energy : +   (kT ) +   (τ  w ) + w F + q H
Dt t
Eq. 5.8
Which can written in scalar form of (r, ϴ, and z) with the aid of cylindrical coordinates. It should be
noted that WF is the work of body forces in rotating frame of reference, F is the body force, while the
subscript r is omitted here. The detailed derivation process of governing equations in rotating frame
of reference can be found in [Schobeiri]88. Alternatively, we can choose more compact form of
integral representation with arbitrary control volume V and differential surface area dA in a relative
frame of reference rotating steadily with angular velocity ω:

dV +  F − G  dA =  S dV where W = ρ , ρu , ρE  and
dW
 v = u − rω
T
V dt V

F = [ρv, ρu  v + p I, ρEv + pu]T G = [0 , τ , τ  v + q]T S = [0 , ρω  u , 0]T


Eq. 5.9
Here F, G and S are respectively, the inviscid flux, viscous flux, and source vectors, and τ, I are stress
and identity tensors respectively. In addition, ρ, u, E, and p are the density, absolute velocity, total
enthalpy, and pressure, respectively and v is the relative velocity. Extended details in available in89.

5.6 Efficiency Effects in Turbomachinery90


In the turbomachinery context a large number of efficiencies are defined such as thermodynamic or
mechanical efficiency. In the sections below the focus is put on the thermodynamic efficiencies. For
a given change of state of a fluid the efficiency is defined as the ratio between actual change in energy
to ideal change in energy in case of expansion or the inverse in case of compression,

actual change in energy


Expansion: η =
ideal change in energy
ideal change in energy
Compression : η =
actual change in energy

88 M. T. Schobeiri, “Turbomachinery: Flow Physics and Dynamic Performance”, Springer, Berlin, 2005.
89“Simulation of unsteady turbomachinery flows using an implicitly coupled onlinear harmonic balance method”,
Proceedings of ASME Turbo Expo 2011, GT2011.
90 Damian Vogt,” Turbomachinery Lecture Notes”, 2007.
103

Eq. 5.10
5.6.1 Isentropic Efficiency
Depending on which process is taken as ideal process efficiencies are referred to as isentropic or
polytrophic efficiencies. In case of an isentropic efficiency the ideal process is represented by an
isentropic change of state from start to end pressure, i.e. the same pressures as for the real process.
This is illustrated in Figure 5.11 for an expansion process by means of an enthalpy-entropy diagram
(h-s diagram). In the above depicted process the changes in total energy are referred to, which is
expressed by indexing the efficiency by “tt”, i.e. “total-to-total”. With the aid of h0 = h + (1/2) c2 where
c is the flow velocity, the total-to-total isentropic efficiency (expansion and compression) is thus given
by
actual change in energy Δh 0 h − h 02
For Expansion : η tt = = = 01
ideal change in energy Δh os h 01 − h 02s
Eq. 5.11
ideal change in energy Δh 0s h 02s − h 01
For Compression : η tt = = =
actual change in energy Δh 0 h 02 − h 01

Figure 5.10 Compression process Figure 5.11 Expansion process

Note: For adiabatic real processes the entropy must always increase during the change of state. Due
to this increase in entropy the real change in energy is smaller than the ideal during expansion. In
other words, you get out less energy from the real process than you could have from an ideal one For
the compression process the increase in entropy signifies that you need to put in more energy to
compress a fluid than you would have in an ideal process Therefore the efficiency is always smaller
or equal to unity The only way to reduce entropy would be to cool a process. However in such case
we do no longer look into adiabatic processes. In certain cases the kinetic energy that is contained in
the fluid (i.e. the amount of energy that is due to the motion) cannot be used at the end of a process.
An example for such a process is the last stage of an energy producing turbine where the kinetic
energy in the exhaust gases is not contributing to the total energy produced. In such case a so-called
total-to-static isentropic efficiency is used, identified by indexing the efficiency by “ts”, i.e. “total-to-
static”. Note that it is necessary to include total and static states in this case. The total-to-total
isentropic efficiency (expansion) is thus given by:
104

−1
actual change in energy h 01 − h 02 Δh 0 1 c 22 
η ts = = = =  +  Eq. 5.12
ideal change in energy h 01−h 2s c 22  η tt 2h 0 
Δh 0s +
2
This relation shows that for values of c2 > 0 the total-to-static efficiency is always smaller than the
total-to-total efficiency. For further detailed aspects of efficiency in turbomachines the readers
should consult with 91-92.

5.7 Case Study 1 – Computation of Heat Transfer in Linear Turbine Cascade


The efficiency of a turbine increases in general with an increase of the temperature of the working
gas which was investigated by [Kalitzin, & Iaccarino]93. This gas temperature may well exceed the
melting temperature of the metal walls. Locally high heat transfer can lead to an excessive
temperature and high thermal stresses in the walls, causing an early fatigue of the high pressure
turbine components. Thus, the design of these components requires accurate evaluation of heat
transfer at the walls (Figure 5.12). The prediction of heat transfer at the end wall and the blade
surface requires simulation of the viscous interaction between the boundary layer approaching the
blade and that developing on the blade itself. Secondary flows, horseshoe type vortices, and strong
turbulence generate complex end wall heat transfer distributions with several local maxima
occurring at the end wall and the blade surface. Accurate prediction of these peaks is crucial for the

Figure 5.12 Linear Turbine Cascade and Computational Domain

91 S.L. Dixon, B.Eng., PH.D., “Fluid Mechanics, Thermodynamics of Turbomachinery”, Senior Fellow at the
University of Liverpool, UK.
92 Damian Vogt, “Efficiencies”, Turbomachinery Lecture Notes, 2007.
93 Kalitzin, G. & Iaccarino G., “End wall heat transfer computations in a transonic turbine cascade”, XVII Congresso

nazionale sulla transmissione delcalore, U.I.T, Ferrara, 1999.


105

design of the turbine cooling system. The objective of the present work is to use this database to
evaluate the influence of turbulence models on the accuracy of heat transfer predictions in
complex three-dimensional flows in turbine geometries. The sensitivity of the heat transfer
coefficient prediction to the turbulence model used is analyzed using two different models: the
Spalart-Allmaras one equation model and Durbin's four equation v2-f model. The use of two different
flow solvers, the NASA research code CFL3D and the commercial package FLUENT©, increases
confidence in the results and allows the elimination of effects related to the numerical discretization
of the equations.
5.7.1 Numerical Methods
The present results have been computed using two different RANS flow solvers: the NASA code
CFL3D and the commercial software Fluent® is a compressible, finite-volume code for multi-block
structured grids. The mean flow fluxes are computed with the Roe flux difference splitting scheme.
Turbulence models are solved segregated from the mean flow in an elimination of effects related to
the numerical discretization of the equations. The CFL3D is a compressible, finite-volume code for
multi-block structured grids. Turbulence models are solved segregated from the mean flow in an
implicit manner using three-factored Approximate Factorization. The v2-f model has been
implemented in this code in an implicit manner. The resulting linear algebraic system is solved with
a three or two-factored Approximate Factorization scheme. Fluent® solves the time-dependent RANS
equations on structured and unstructured meshes using a control-volume-based technique; the
diffusion terms are discretized using a second-order central-difference scheme while a second-order
upwind scheme is employed for the convective terms. An Euler implicit discretization in time is used
in combination with a Newton-type linearization of the fluxes. The resulting linear system is solved
using a point Gauss-Seidel scheme in conjunction with an algebraic multi-grid method. The additional
equations for the turbulent quantities are solved in a segregated fashion using a 1st or 2nd order
upwind discretization scheme with explicit boundary conditions.
5.7.2 Mesh Generation
The large scale linear cascade investigated in the experiments consists of twelve blades with an axial
chord of 10.7 cm. A part of the cascade is shown in Figure 5.13 A. The high blade count of the
cascade ensures good periodicity. This allows us to consider only one blade and only the region
between end wall and symmetry plane in the computations. The actual computational domain is

Shock
Reflection

Figure 5.13 A) Default mesh B) Refine mesh


106

shown in Figure 5.13 B. The block boundaries of the structured 3-block mesh and the boundary
conditions used are highlighted in the Figure 5.12. An O-mesh topology around the blade has been
chosen to ensure a high quality mesh near the blade surface. The two-dimensional mesh consisting
of 48x192 cells has been generated through simple geometric interpolation. After generating the
outer boundary as an arbitrary line between two blades and distributing lines connecting the outer
boundary with the blade, O-lines have been interpolated using a stretching function. The three-
dimensional mesh has been generated by copying the described 2D grid in the span-wise direction
and clustering the grid points at the end-wall. Two meshes, mesh A and mesh B, have been generated
with 40 and 52 cells span-wise, respectively. All block dimensions have been chosen to contain
factors of the power 2 to exploit multi-grid. The mesh has been transformed into an unstructured
mesh for the flow computations with Fluent©. The multi-block decomposition disappears for an
unstructured solver. The height of the first cell above the wall has an average y+ value of about 1. The
height has been adjusted after initial computation.
5.7.3 Heat Transfer Results for 2D & 3D
In the simulation of three-dimensional flow, the computational grid is often a compromise between
a desired resolution and computational accord ability. In two dimensions, however, it is easier to
carry out a complete grid sensitivity study. With this objective in mind, the flow in the symmetry
plane has been computed in a two-dimensional plane. The structured grid or default mesh, for this
report is shown in Figure 5.13(A). It is the same used at each span wise location in the three-
dimensional calculations. It contains 11008 cells. The unstructured grid, shown in Figure 5.13(B)
is obtained through successive refinement in regions with high pressure gradients and large strain
rates like shock waves, boundary layers, and wakes. This mesh contains 71326 cells. The Mach
contours plotted for both grids show a very complex shock wave pattern in the wake of the blades.
The accelerating flow within the passage generates an oblique shock wave on the pressure side of a
blade (see red circles in Figure 5.13). This shock is reflected on the suction side of the successive
blade. It then interacts with the viscous wake of the blade from which it originated. Partly due to
reflection in pressure BC. Somehow the new development by ANSYS© claims that Average Pressure
Specification at Pressure
Boundary which allows the f=0 f = 0.5 (Old)
exit pressure to vary across
the boundary, but
maintains an average
equivalent to the specified
exit pressure value94. It also
claims that it is less
reflective than previous
version with improved
results. The Pressure
blending factor ‘f’ (default
value 0.0) may need to
change f > 0.0 in cases
where stability is degraded.
For f = 0 recovers the fully Figure 5.14 Average Pressure Specification at pressure boundary
averaged pressure, and f = 1
recovers the specified pressure. The results of this improvement displayed in Figure 5.14.
A second shock wave is generated on the suction side near the trailing edge. The default mesh does
not resolve the shock wave the wake. Only the two shocks at the trailing edge are clearly visible. The

94 Ansys Fluent© 16.0 Preview 4.


107

heat transfer at the wall depends significantly on the thermal conductivity of the fluid. The effect of
using a constant thermal conductivity at reference temperature is demonstrated with the FLUENT©
results reported in the same figure. The overall Stanton number is under-predicted. This explains the
difference observed between the FLUENT and CFL3D Stanton number distributions at the end wall
reported. It has to be noted that the constant thermal conductivity is the default option in. The
pressure distributions on blade and end-wall are not very sensitive to the grid resolution and inflow
profile for the case considered. Both flow solvers predicted a reasonable agreement with the
experiment as reported in. We note, however, that the pressure distribution on the blade and the
shock structure is sensitive to the treatment of the periodic boundary since it is located relatively
close to the blade surface. In this paper we will focus primarily on the analysis of the heat transfer
distribution, on the dependence of the Stanton number distribution on inflow profile and grid
resolution.
5.7.4 Experimental Data
The experimental data for the end-wall show some interesting features that will help to differentiate
the predictive capabilities of the models tested (Figure 5.15). The horseshoe vortex generated by
the rolling up of the incoming boundary layer enhances the wall heat transfer, and its structure is
clearly visible in the higher Stanton number (Region A). A second distinct heat transfer peak is
measured near the stagnation point (Region A). Within the passage, four additional interesting
features are present: the first is a localized peak in the Stanton number related to the impingement
of the suction-side leg of the horseshoe vortex on the blade surface (Region B). The second feature is
the presence of a shock wave on the pressure side near the trailing edge that increases the heat
transfer on the end wall (Region C). Third, there is a gradual increase of heat transfer at the end wall
which is related to the acceleration of the fluid in the passage (Region D). And finally, the presence of
a corner vortex on the suction side of the blade (Region E) is indicated in experiments by a low heat
transfer region. In the wake, a very sharp peak in the Stanton number is measured just downstream

Figure 5.15 Stanton Number Distribution on End-Wall


108

of the trailing edge (Region C). The numerical predictions of the Stanton number show most of the
features observed in the experiments but, in general, fail to predict the quantitative heat transfer on
the end wall correctly.
5.7.5 Effects of Turbulence
The increased heat transfer beneath the horseshoe vortex is captured by both turbulence models.
The S-A model seems to spread this high Stanton number region and shift it towards the suction side.
Spreading of the horseshoe vortex is related to the turbulence generation in the vortex shear layer.
The v2-f model tends to produce a thinner vortex. The secondary peak on the suction side (Region
B), which is related to the stagnation of high temperature fluid convected by the horseshoe vortex, is
predicted by both models. The v2-f model predicts a higher value for the Stanton number. The SA
model predicts slightly larger values for the gradual increase in heat transfer within the passage
(Region D). The trailing edge peak (Region C) and the low heat transfer region on the suction side of
the blade (Region E) are reproduced by both models. A quantitative comparison of the heat transfer
on the blade surface is shown for three stations in Figure 5.16 for the v2-f and SA model,
respectively.

Figure 5.16 Stanton number Distribution on Blade Surface for 2D Grid

The heat transfer in the stagnation region, the location where span is 0, is accurately predicted by
both models at 25% and 50% span (solid line). Both stations are located outside of the incoming
boundary layer specified at the inlet. The station at 10% span, however, is located well inside of this
boundary layer, and both models over-predict the heat transfer here by 25%. The higher heat
transfer indicates that the turbulence intensity is too high at this location. This observation is
supported by a computation in which the turbulence levels inside the end wall boundary layer have
been reduced by setting the turbulence quantities at the inlet to a uniform value corresponding to
25% turbulent intensity (dotted line). This lowers the Stanton number in the stagnation region to
the value measured in the experiments. In addition, it delays the transition on the upper surface of
the blade. The SA model shows a large sensitivity to the reduced boundary layer turbulence across
the entire span on the pressure side. The heat transfer on the pressure side of the blade is consistently
under-predicted at each station by both models. The same has been observed for the 2D computation
shown in Figure 5.16. At this stage it is not clear whether this is due to the specification of the inlet
109

conditions or the turbulence model95.

5.8 Case Study 2 - Using Shock Control Bumps To Improve Transonic Compressor
Blade Performance96
Citation : John, A., Qin, N., and Shahpar, S. (March 2, 2019). "Using Shock Control Bumps to Improve
Transonic Fan/Compressor Blade Performance." ASME. J. Turbomach. August 2019; 141(8):
081003. https://doi.org/10.1115/1.4042891
Shock control bumps can help to delay and weaken shocks, reducing loss generation and shock-
induced separation and delaying stall inception for transonic turbomachinery components, as
described by [John et al.]97. The use of shock control bumps on turbomachinery blades is investigated
here for the first time using 3D analysis. The aerodynamic optimization of a modern research fan
blade and a highly loaded compressor blade are carried out using shock control bumps to improve
their performance. Both the efficiency and stall margin of transonic fan and compressor blades may
be increased through the addition of shock control bumps to the geometry. It is shown how shock
induced separation can be delayed and reduced for both cases. A significant efficiency improvement
is shown for the compressor blade across its characteristic, and the stall margin of the fan blade is
increased by designing bumps that reduce shock-induced separation near to stall. Adjoint surface

Figure 5.17 Contours Of Casing Static Pressure Beneath A High-Speed Rotor (550 M/S
Tip Speed) With Pronounced Negative Camber. From Prince – Courtesy of [Prince]

95 Kalitzin, G. & Iaccarino G., “Computation of heat transfer in a linear turbine cascade”, Center for turbulence
Research Annual Research Briefs, 1999.
96 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic

Fan/Compressor Blade Performance”, GT2018-77065.


97 See Previous.
110

sensitivities are used to highlight the critical regions of the blade geometries, and it is shown how
adding bumps in these regions improves blade performance. Finally, the performance of the
optimized geometries at conditions away from where they are designed is analyzed in detail.
5.8.1 Introduction and Motivation
Shocks are a major source of loss for transonic fans and compressors. They cause entropy generation,
boundary layer thickening and shock induced separation. The impingement of the shock on the blade
suction surface (and the resulting, strong, adverse pressure gradient) can cause the boundary layer
to detach, leading to larger blade wakes, reduced efficiency, lower blade stability and reduced stall
margin. Any method that can be used to alleviate shock strength (and the associated negative effects)
therefore has the potential to significantly improve transonic fan/compressor performance.
5.8.2 Shock Control for Turbomachinery & Literature Survey
Relatively little work on designing geometries directly to weaken the shock waves in transonic
turbomachinery components can be found in the literature, though it has been known for some time
that reducing the pre-shock Mach number of transonic compressors can improve their efficiency98.
It was clear to transonic compressor designers in the 70s and 80s that shock strength was increased
by the amount of convex curvature on the suction side between the leading edge and the shock99.
Nearly flat suction surfaces that minimized the expansion were therefore favored, with the next step
to try designs with concave curvature (often referred to as negative camber). Geometries with
negative camber result in gradual compression along the suction surface which weakens the shock.
The concave curvature of the blade surface and the reduction of flow area in the flow direction leads
to a deceleration of the supersonic flow through compression waves, and therefore a weaker shock.
[Prince]100 designed a rotor with pronounced negative camber (see Figure 5.17). This lead to a rise
in static pressure along the suction surface prior to the shock as intended, but the resulting efficiency
was disappointing due to the strong shock on the pressure surface. [Ginder and Calvert]101 had more
success in designing a rotor with negative camber. With negative camber, the Mach number ahead of
the shock was reduced to 1.4 (compared to 1.5 for the traditionally designed blade) which drastically
reduced the amount of boundary layer separation and loss.
Recently, it was demonstrated by [John et al.]102 how the freeform shaping of a compressor blade can
improve blade efficiency by delaying and weakening the shock and reducing separation. The flexible
parameterization method used allowed an s-shaped design to be generated that included a pre-
compression geometry around mid-span. This s-shaped, pre-compression geometry is similar to the
negative camber designs described above. The effect of the pre-compression geometry on the shock
and separation is described in Figure 5.18. The current work proposes the use of shock control
bumps as an alternative method to reduce shock related loss to those described above. Shock control
bumps have the benefit that relatively small modifications to the original geometry are required to
achieve the desired effect.

98 Ginder, R., and Calvert, W., 1987. “The design of an advanced civil fan rotor”. Journal of turbomachinery,
109(3), pp. 340–345.
99 Cumpsty, N. A., 1989. Compressor aerodynamics. Longman Scientific & Technical.
100 Prince, D. C., 1980. “Three-dimensional shock structures for transonic/supersonic compressor rotors”. Journal

of Aircraft, 17(1), pp. 28–37.


101 Ginder, R., and Calvert, W., 1987. “The design of an advanced civil fan rotor”. Journal of turbomachinery,

109(3), pp. 340–345.


102 John, A., Shahpar, S., and Qin, N., 2017. “Novel compressor blade shaping through a free-form method”. Journal

of Turbomachinery, 139(8), p. 081002.


111

Figure 5.18 Schematic of Shock Structures (A) Datum, (B) S-Shaped Design. Courtesy of [John et al.]

5.8.3 Shock Control Bumps


Shock control bumps are bumps added to aerodynamic surfaces to alter the behavior of the shock
and improve aerodynamic performance. One of the earliest examples of 2D shock control bump usage
is in the design of the dromedary foil in the 1970s. This was a modified supercritical airfoil with a
bump added in an attempt to increase its drag-divergence Mach number. The ’hump’ was shown to
weaken the shock wave when implemented in the right position, acting as a localized precompression
ramp. This also demonstrated the importance of shock control bump positioning, as, if the bump was
misplaced, an increase in wave drag was seen. [Ashill et al.]103 found for a 2D airfoil a significant
reduction in drag could be achieved via the correct application of a shock control bump, however
when the shock position changed severe drag penalties were incurred due to secondary shocks and
separation being produced. [Drela and Giles]104 carried out numerical studies into shock control in
1987, describing the behavior of shock-induced separation. [Sommerer et al.]105 optimized shock
control bumps at various Mach numbers. They concluded that the bump height, width and position
of the bump peak are the key parameters. [Collins et al.]106 tested shock control bumps in a wind
tunnel, and analyzed the performance of shock control bumps at off-design conditions.
The EUROSHOCK II project began in 1996 and concluded that shock control bumps had the most
potential out of a range of shock control devices tested. A large amount of research was carried out
into shock control bumps, with both 2D and 3D analysis, although no optimization was undertaken.
[Qinet al.]107 first proposed 3D shock control bumps with a finite width, allowing additional design
complexity. They showed that 3D bump configurations were more robust than 2D bump designs

103 Ashill, P., and Fulker, J., 1992. “92-01-022 a novel technique for controlling shock strength of laminar-flow
aerofoil sections”. DGLR BERICHT, pp. 175–175.
104 Drela, M., and Giles, M. B., 1987. “Viscous-inviscid analysis of transonic and low Reynolds number airfoils”.

AIAA journal, 25(10), pp. 1347–1355.


105 Sommerer, A., Lutz, T., and Wagner, S., 2000. “Design of adaptive transonic airfoils by means of numerical

optimisation”. In Proceedings of ECCOMAS, 2000, Barcelona.


106 Colliss, S. P., Babinsky, H., N¨ubler, K., and Lutz, T., 2016. “Vortical structures on three-dimensional shock

control bumps”. AIAA Journal, pp. 2338–2350.


107 Qin, N., Wong, W., and Le Moigne, A., 2008. “Three dimensional contour bumps for transonic wing drag

reduction”. Proceedings of the Institution of Mechanical Engineers, Part G: Journal of Aerospace Engineering,
222(5), pp. 619–629.
112

(where a 2D bump is extended continuously along the span).


The only use of a shock control bump on turbomachinery blades found in the literature is by
[Mazaheri and Khatibirad]108, who tested a 2D shock control bump on a (mid-span) section of the
NASA rotor 67 geometry. They added a bump modelled using the Hicks-Henne function. It was shown
how the interaction of the bump with the original wave structure resulted in a more desirable
pressure gradient, with a weaker compression wave fan and a more isentropic compression field.
The bump design was optimized and was shown to reduce the separation area at an off-design
condition. They describe how this may have the potential to improve the stall properties of the blade
section. Two optimizations were carried out, one at the design condition and another at 4% higher
rotational speed. Optimal bumps were produced for each condition, with an increase in efficiency of
0.67% for the on-design case and 2.9% in the off design case reported. The optimized geometry for
the design condition is shown in Figure 5.19.

Figure 5.19 Datum Geometry and Optimized Shock Control Bumps on The Mid- Section of Nasa Rotor
67- From Mazaheri et al..

The work by [Mazaheri & Khatibirad]109 demonstrated the benefit that bumps may provide at both
on and off-design conditions, and their potential to improve stall margin. The simplified 2D analysis
lacks accuracy however as the complex behavior of radial and separated flow cannot be predicted.
For a thorough understanding of the potential for the use of shock control bumps, 3D analysis and
the design of 3D bumps is needed to truly assess their effect.
5.8.4 Test Case - NASA Rotor 37
The case studied here is NASA Rotor 37. This has a very strong shock wave (with a relative tip Mach
number of nearly 1.5) which causes large separation, decreasing the blade efficiency. It is a well-
documented case, having been extensively tested and simulated as part of a turbomachinery
validation study. It is a transonic rotor with inlet hub-to-tip ratio 0.7, blade aspect ratio 1.19, rotor
tip relative inlet Mach number 1.48 and rotor tip solidity 1.29. It has historically been a challenge for
CFD simulation. The very high pressure ratio, strong shock wave-boundary layer interaction, large
tip leakage vortex and highly separated flow mean that it poses challenges for turbomachinery

108 Mazaheri, K., and Khatibirad, S., 2017. “Using a shock control bump to improve the performance of an axial
compressor blade section”. Shock Waves, 27(2), pp. 299–312.
109 See Previous.
113

solvers. Rotor 37 has been the subject of review articles that highlight the complexity of matching
experimental and computational measurements and the associated uncertainties.
The CFD setup is shown in Figure 5.20. At the inlet, a radial distribution of total pressure and
temperature (based on the original experimental values) is specified. The inlet turbulence intensity
is 1%. At the outlet, a value for circumferentially mixed-out and radially mean-mass capacity (non-
dimensional mass flow) is used. Periodic boundaries are used to represent full annulus flow.

Figure 5.20 The R37 CFD Domain Used – Courtesy of [John et al.]

Stationary walls are treated as adiabatic viscous walls and the rotational speed of the non-stationary
portions of the domain is 1800.01rads􀀀1, as specified in the experiment. Rolls-Royce CFD solver
Hydra is used for all of the simulations presented here, using the Spalart-Allmaras turbulence model
(fully turbulent). The 4.27 million cell mesh is generated by PADRAM, has y+ of the order of one on
all surfaces with 30 cells in the tip gap.
5.8.5 Validation
As previously alluded to, many studies have struggled when matching simulations of Rotor 37 to the
experiment. A wide range of work has been undertaken to investigate the discrepancy found between
simulation and experiment, with the primary work being the 1994 ASME/IGTI blind test case study
in which a range of codes were used to simulate the rotor, with no knowledge of the experimental
values. A large variation was seen between the different predictions, prompting analysis by
114

[Denton]110. Recent work has also been carried out by [Chima]111 and [Hah]112. The differences are
usually attributed to uncertainty in the experimental measurements, the lack of real geometry in the
simulations (e.g. the upstream hub cavity is usually missing) and also the difficulty in fully resolving
the complex flows. The pressure ratio agreement is reasonable across the characteristic, but the
efficiency prediction is about 2% below the experimental value at the design point (98% choke). This
matches the trend of previous results, where the better the PR prediction, the worse the efficiency
match. This ’trade-off’ has been seen in a range of previous simulations. Figure 5.21 gives the radial
profiles of total PR and efficiency at 98% of simulated choke compared to the experimental values at
98% experimental choke. The radial trends have been captured fairly well, although there is an offset
from the experiment for both. The choke mass flow found in the simulations was 20.91kg/s, matching
quite closely the experimental of 20.93kg/s.

Figure 5.21 Radial Profiles Vs Experimental Data – Courtesy of [John et al.]

5.8.6 Flow Field For the Datum Case


Figure 5.22 shows the flow features of the datum NASA Rotor37 at design point. It can be seen how
the strong shock of Rotor 37 causes complex shock-boundary layer interaction and a large shock-
induced separation (this can be seen by the thickening of the boundary layer and wake shown in
Figure 5.22-b and the orange contour of zero axial velocity in Figure 5.22-a. At the point where
the shock impinges on the suction surface, its interaction with the boundary layer causes it to

110 Denton, J., 1997. “Lessons from rotor 37”. Journal of Thermal Science, 6(1), pp. 1–13.
111 Chima, R., 2009. “Swift code assessment for two similar transonic compressors”. In 47th AIAA Aerospace
Sciences Meeting including The New Horizons Forum and Aerospace Exposition, p. 1058.
112 Hah, C., 2009. “Large eddy simulation of transonic flow field in NASA rotor 37”. 47th AIAA Aerospace

Sciences Meeting including The New Horizons Forum and Aerospace Exposition, p. 1061.
115

separate and a large wake forms. It is at this design point that Rotor 37 will be optimized, as a
reduction in this separation could significantly increase blade efficiency.
5.8.7 Validation
Due to experimental data for this geometry not being available, simulation validation was carried out
using a similar fan blade geometry that has experimental data available. The related blade has very
similar performance parameters, and the simulation set up is identical. The results are given here. A
comparison of the simulations of this related blade against experimental data can be seen in 113. Both
the pressure ratio and efficiency curves match the experimental data well, though there is a slight
offset to the overall values and stall margin. The radial curves show good comparison to experimental
data, although the radial variation in efficiency is underpredicted compared to the experiment.
Overall, the simulation compares well, lying within 1% across the range of flow rates.

Figure 5.22 (a) 3d Separation (Orange) On The R37 Geometry (Flow Right To Left), (b) Rel. Mach
No. Contour At 60% Span – Courtesy of [John et al.]

5.8.8 Blade Flow Features


To understand the behavior of this blade design and select a point at which to optimize the geometry,
the flow behavior for a range of flow rates was studied. Figure 5.23 shows the flow features of the
blade design as the flow rate is varied (see [John et al.]114. Point A is stalled. For proprietary reasons
the whole RR-FAN blade geometry cannot be shown, hence, flow behavior in just the region of
interest is shown in the following figures. The shock position on the blade surface moves towards the
LE as the operating point moves to the left on the characteristic. As the pressure ratio increases and
flow rate becomes lower, the strength of the shock increases and separation is caused towards stall.
It is this separation (highlighted in orange) that contributes to the full stall of the blade. It can be seen
that the shock-induced separation increases in magnitude and radial extent as the flow rate is
lowered until full separation eventually occurs. These near-stall operating points are a promising

113 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic
Fan/Compressor Blade Performance”, GT2018-77065.
114 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic

Fan/Compressor Blade Performance”, GT2018-77065.


116

area to investigate the benefit of shock control bumps. It is the shock-induced separation that is
responsible for limiting the operating range of the blade, and if this separation can be reduced then
it is expected that this will extend the stable working range of this fan.

Figure 5.23 Shock Region Flow Features For RR-Fan At Points A) A, B) B, C) C, D) D, E) E, F) F. Flow
Direction – Courtesy of [John et al.]
RIGHT TO LEFT.
5.8.9 Adjoint Sensitivity Analysis
Adjoint sensitivity analysis is a useful tool that can be used to provide information on the sensitivity
of an objective function to changes in the geometry. Here, the adjoint sensitivity used is the
sensitivity of efficiency (as a percentage) to surface deformation (in mm) normal to the surface. This
can be used to inform which regions of the blade will have the greatest impact when modified, and
are therefore most important to control during an optimization. Hydra Adjoint115 is used to provide
the blade surface sensitivities: A primal Hydra simulation is first used to provide the flow solution,
followed by Hydra adjoint which calculates the flow-adjoint sensitivity and provides the sensitivity
of the objective function to changes in the flow. Once these two relatively expensive simulations are
completed, the mesh sensitivities are then mapped onto the surface. This finds the relationship
between changes in the flow to changes in the blade surface mesh. Combining these provides the
sensitivity (gradient) of the objective function (efficiency) to perturbations of the blade surface. The
adjoint surface sensitivity analysis for Rotor 37 at design point and RR-FAN at point D. It can be seen

115 Duta, M. C., Shahpar, S., and Giles, M. B., 2007. “Turbomachinery design optimization using automatic
differentiated adjoint code”. ASME Turbo Expo 2007: Power for Land, Sea, and Air, American Society of
Mechanical Engineers, pp. 1435–1444.
117

that the most sensitive regions of both geometries are focused around the shock on the suction
surface. This indicates that geometry changes in this region will have a significant impact on the blade
efficiency, and therefore if shock control bumps are applied here some benefit should be found. For
complete details, please consult the [John et al.]116
5.8.10 Shock Bump Parameterization & Optimization
The CST (Class Shape Transformation) method is used in this work to define the bump geometries.
The CST method uses Bernstein polynomials to create smooth (second derivative continuous)
contour bumps. For this project 3rd order CST bumps are used, constructed from four Bernstein
polynomials.
Controlling the
weighting (amplitude)
of these polynomials
modifies the bump
height and asymmetry.
Figure 5.24 shows
how the bump
geometry (solid black
line) to be added to the
blade surface is a sum of
the four Bernstein
polynomials (colored
dashed lines). The CST
bump parameterization
provides a high degree
of flexibility, enabling
the generation of Figure 5.24 Example 2d CST Bump (Solid Line) And The Four Polynomials
Used To Construct It (Dashed Lines) – Courtesy of [John et al.]
smooth, asymmetric
bumps in 2D and 3D.
The CST bump parameterization
technique was implemented inside of
the PADRAM geometry and meshing
software. The technique modifies each
2D radial section of the blade geometry,
adding a bump. The properties of these
2D bumps are smoothly interpolated in
the radial direction from control
sections. The resulting geometry is
controlled by the bump start and end
positions, the four Bernstein
polynomial amplitudes and the span-
wise distribution. This allows 3D
variation of the bumps in the radial
direction. Both continuous (where Figure 5.25 a) Example Individual Bump Geometry
bump amplitudes are smoothly And B) Example Continuous Bump Geometry – Courtesy of
interpolated radially) and individual [John et al.]
(where the bump amplitude returns to

116Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic
Fan/Compressor Blade Performance”, GT2018-77065.
118

zero periodically in the radial direction) CST bumps were tested. Examples of the blade with
individual and continuous bumps added is shown in Figure 5.25.
During this work, a study was carried out (not detailed here for brevity) to compare the benefit of
using individual bumps (where a series of discrete bumps is added to the datum geometry in the
radial direction) with a continuous bump (note continuous bumps are still ’3D’ and their shape,
position and amplitude can vary in the radial direction). It was concluded that, for these cases, the
individual bumps needed to have greater amplitude than the continuous bumps to offer the same
benefit, leading to increased separation downstream of the bump position. The continuous bumps
tested offered greater benefit, and therefore only results using the ’continuous’ bump geometry
approach are presented here.
5.8.11 Optimization Method
In this work the Multi-point
Approximation Method (MAM) is
used for the optimization studies. It is
a gradient based method that uses
localized Design of Experiments (DoE)
and trust regions to efficiently search
through the design space. When using
MAM, an initial generation of
simulations (chosen by DoE) is carried
out around the start point. A response
surface is constructed for this region
and the sub-optimal point found. The
search is then moved to this point,
where a new generation is constructed Figure 5.26 Spanwise Slice of the Datum and Optimized
and the process repeated until the R37 Geometries At 60% Span – Courtesy of [John et al.]
search converges on the optimal
design. The MAM method has been shown to be an efficient and consistent approach for a wide range
of highly constrained optimization problems, working successfully for design spaces made up of
hundreds of parameters.
5.8.12 Rotor 37 Bump Optimization
For the Rotor 37 optimization, the bump geometry was
controlled at 5 radial heights (to allow radial variation of
the parameters) with the geometry smoothly interpolated
between the control stations using a cubic B-spline.
Towards the tip the bump placement and movement range
are increased in chord-wise position as the shock is sat
further downstream at the tip. The initial design used at
the start of the optimization process had bumps positioned
with approximately 60% of the bump downstream of the
datum shock, as is known to be beneficial from previous
work. The objective function for the optimization was
blade efficiency and the simulations were carried out at
98% simulated choke. The optimizations were carried out
on the Rolls-Royce CFMS cluster using the MAM method.
The geometry of the optimized shock bump can be seen in
Figure 5.26. A slice at 60% span is shown. The 3D Figure 5.27 Optimized R37 Bump
geometry compared to the datum is shown in Figure 5.27. (Blue) Added To The Datum Blade
Geometry (Grey)
The bump applied to the datum geometry varies radially,
119

with the maximum bump amplitude and width localized between 40 and 60% span. This makes sense
as the strongest shock location, largest separation and maximum adjoint sensitivity occur around
mid-span for Rotor 37, and therefore greater shock control is needed in this region. The resulting
variation from hub to tip of the geometry demonstrates the benefit provided by optimizing the
geometry. Without optimization it would be difficult to manually specify the bump position, width,
amplitude and asymmetry, which would result in reduced benefit.
5.8.13 Analysis of the R37 Optimized Bump Design
The flow features for the resulting, optimized, continuous bump design is compared to the datum in
Figure 5.28. The datum shock position is shown via a white line on the optimized geometry. It can
be seen how the use of bumps has delayed the shock. The reduction in separation for the optimized
design can be seen in Figure 5.30. The delay of the shock position has reduced the separation
initiation point and the volume of separated flow. The performance of this geometry is compared to
the best
individual PR Delta PR / % Efficiency / % Delta efficiency / %
bumps Datum 2.05 - 85.45 -
geometry
Individual 2.06 0.51 86.21 0.76
(not
Cont. 2.08 1.2 86.93 1.48
described
in detail Table 5.1 Rotor 37 Optimized Bump Performance Comparison – Courtesy of [John et al.]
here) and
the datum
in Table 5.1. It can be seen that the efficiency benefit is greatest for the continuous bump design.
The efficiency is increased by 1.48%, while the pressure ratio is also increased. A summary of

Figure 5.28 Datum (Left) And Optimized (Right) Rotor 37 Static Pressure Contours. Flow Direction
Right To Left – Courtesy of [John et al.]

previous optimization results for Rotor 37 by various researchers is given by [John et al.]117. The
maximum efficiency benefit achieved by those studies was around 1.7-1.9% (without decreasing PR).
These optimizations were able to modify parameters such as blade camber, thickness, lean and
sweep though, so had greater design flexibility than the current shaping approach. This shows that

117 John, A., Shahpar, S., and


Qin, N., 2017. “Novel compressor blade shaping through a free-form method”. Journal
of Turbomachinery, 139(8).
120

the efficiency benefit provided


through the application of shock
control bumps is significant,
considering the only geometry
change is the addition of bumps.
Figure 5.29 shows the passage
flow for the datum and optimized
geometries at 50% span. The
effect of the bump delaying the
shock can be seen, with the datum
shock position shown by the black
line. The shock has been delayed
by over 12% chord at this height.
Just upstream of the shock the
Mach number contour is lower,
Figure 5.29 Datum (Left) And R37 Optimized (Right)
suggesting pre-compression has Flow Features At 50% Span – Courtesy of [John et al.]
occurred. The boundary layer
separation that forms the wake,
highlighted by the dark blue, low velocity region, has reduced in width by 26% at the trailing edge
for the optimized design. Figure 5.31 shows the datum and optimized lift plots. It can be seen how
the shock has been delayed. The Cp increases just upstream of the shock, showing that the bump has
carried out pre-compression. The jump in pressure across the shock is also lower for the optimized
design than for the datum, indicating it has been weakened. Because the shock is delayed, it has
become swallowed by the passage, causing an acceleration near to the leading edge on the blade
pressure surface. This can be seen in the lower surface spike on the lift plot.

Figure 5.30 Datum (Left) And Optimized (Right) Rotor 37 Separated Flow Contours (Orange). Flow
Direction Right To Left.

5.8.14 Performance Across the Characteristic for R37


The off-design performance is a key feature of blade aerodynamics. The characteristics for the datum
and optimized designs are shown in Figure 5.32. An efficiency and pressure ratio increase has been
achieved across the characteristic. The choke mass flow does not appear affected, although it is
possible that the choke margin has been modified at other rotor speeds due to the throat area being
reduced by the bump. The simulation results suggest a reduction in stall margin for the optimized
121

design. This is due to the shock


bump being mis-placed at
conditions away from where it was
designed, leading to increased
separation and thus reduced stall
margin. This section has
demonstrated the benefit that can
be achieved by applying shock
control bumps to a compressor
blade without modifying the entire
blade geometry. This shows that a
significant benefit is possible
through geometry modifications
via bumps just in the shock region.
5.8.15 Conclusion
This work has demonstrated how
shock control bumps can be used
Figure 5.31 Lift Plots For The Datum and Optimized
to improve the performance of Geometries at 60% Span – Courtesy of [John et al.]
transonic fan/compressor blades.
Blade geometries that incorporate
shock control bumps have the ability to reduce shock loss and reduce/eliminate shock-induced
separation and increase both efficiency and stall margin. Shock control bumps have the benefit that
only small modifications to the blade geometry are required to achieve these improvements,
compared to the large changes required by blade designs that make use of negative camber or similar
shock control approaches.
It has been demonstrated
that both the efficiency
and pressure ratio of a
highly loaded compressor
blade can be increased
across a range of flow rates
by delaying the shock and
significantly reducing the
separation and wake. For a
modern fan blade the
optimized bump design
eliminated the majority of
separation, reduced the
thickness of the wake and
extended the stall margin.
For further and complete
info, please consult the [
John et al.]118.

Figure 5.32 R37 Optimized Characteristic Vs Datum – Courtesy of [John


et al.]

118 John, A., Shahpar, S., and


Qin, N., 2017. “Novel compressor blade shaping through a free-form method”. Journal
of Turbomachinery, 139(8).
122
123

6 Complex Flow in Turbomachinery


6.1 Key Features of Transonic Fan (Turbine) Field
These features include highly 3D flow fields, complex shock systems, and strong interactions between
the shock, boundary layer, and secondary flows (like the tip-leakage vortex). The goal is to provide a
basic understanding so that proper assessment of the chosen numerical approach can be performed.
As suggested by Figure 6.1, the flow fields of fan designs are complex and highly three-dimensional,
and almost always unsteady. The flow-path hub contour shown in Figure 6.1 suggests significant
radial velocity components, especially at the fan entrance and strong interactions between the shock,

Figure 6.1 Complex Flow Phenomena Compressors

boundary layer, and secondary flows (like the tip-leakage vortex). Secondary flows and their
interactions with other phenomena are another major source of flow complexity. Indeed, Denton and
Dawes, 1999, suggest the prediction of blade surface and end-wall corner separations to be one of
the most challenging tasks of 3D, viscous solvers, largely due to the obvious dependence on
turbulence model. Additionally, the use of blade twist, sweep (viewed from the meridional plane)
and lean (observed looking axially through the machine) contributes to the 3D flow effects.
A significant consideration in the design of transonic fan blades is the control of shock location and
strength to minimize aerodynamic losses without limiting flow. Custom-tailored airfoil shapes are
required to “minimize shock losses and to provide desired radial flow components. Figure 6.2 shows
features of the tip section geometry typical of a transonic fan. The shape of the suction surface is key
as it:
• Influences the Mach number just ahead of the leading edge passage shock, and
• Sets the maximum flow rate.
124

As noted by Wisler, 1987, the cascade passage area distribution is chosen to provide larger-than-
critical area ratios; thus, maximum flow is determined by the first captured Mach wave, location
determined by the forward suction surface (induction surface). This maximum flow condition is often
referred to as leading edge choke, or in cascade parlance, “unique incidence” (note that “unique”
incidence is really a misnomer; here, “choking” incidence will be used). The flow induction surface
and fan operating condition (incoming relative Mach number at the airfoil leading edge) set the
average Mach number just ahead of the leading edge passage shock. A “traditional” convex suction
surface results in a series of Prandtl-Meyer expansion waves as the flow accelerates around the
leading edge. Increasing the average suction surface angle (relative to the incoming flow) ahead of
the shock reduces the average Mach number, and presumably reduces the shock losses. Common for
modern transonic fan tip sections is a concave induction surface, the so-called “pre-compression”
airfoil. As indicated in previous chapter, there are four major area of research going on in
turbomachinery, namely: Unsteady Flow, Film cooling, Turbulence and 3D Flow. We start with the
unsteadiness first.

Figure 6.2 Fan Tip Section Geometry

6.2 Sources of Unsteadiness in Turbomachinery


Turbomachinery flows are among the most complex flows encountered in fluid dynamic practice
(Lakshminarayana,)119. The internal flows within a blade passage of turbomachinery are strongly
three dimensional, viscous flows which may include laminar flow, turbulent flow and transitional
flow. Moreover, they are fully unsteady due to the interactions between blade rows in a stage or
multistage machine. There also exist secondary flows including the flows due to passage vortices in
the end-wall range, radial flow near blade surfaces, and tip leakage flow and leakage vortex, shock
and shock boundary layer interaction in high speed conditions, wakes flows, even some specific
flows, for instance film cooling flows nearby the cooling holes. The complexity is mainly reflected in
the following areas:

119Lakshminarayana, B. “An assessment of computational fluid dynamic techniques in the analysis and design of
turbomachinery”, the 1990 freeman scholar lecture, J. Fluids Engineering Vol. 113(No. 3): 315-352, 1991.
125

1. Various forms of secondary flow caused by viscosity and complex geometry, which is
dominated by vortex flows: passage, leakage, corner, trailing, horseshoe and scraping
vortices, etc. These form three- dimensional and rotational nature of the flow.
2. Inherent unsteadiness (see below) due to the relative motion of rotor and stator blade
rows in a multi stage environment.
3. The flow pattern in the near-wall region includes: laminar, transitional and turbulent
flows; besides separated flows are often exist.
4. The flow may be incompressible, subsonic, transonic or supersonic; some
turbomachinery flows include all these flow regimes.
5. Due to the limitation of flow space, there are strong interactions of the solid wall surfaces
with above complicated phenomena. Besides, in gas turbines, the use of cooling gas makes
the flow more complex.

A good understanding of the unsteady flow in turbomachinery is necessary for advanced design as it
shown in Figure 6.3 with broad spectrum. According to Greitzer120, the unsteady flow in
turbomachinery can be classified into two groups: inherent unsteadiness and conditional
unsteadiness. The conditional unsteadiness is mainly caused by the sudden changes of the working
condition. For example when turbomachinery is working on the start stage, acceleration stage or off-
design condition, the fluctuation of working condition might lead to the unsteady rotating stall,
surge, flutter and flow distortion of turbomachines. Sometimes, the distortion of inlet flow or the
asymmetric outlet condition of vector nozzle also might lead to the unsteadiness. The inherent
unsteadiness is mainly due to the relative motion and interaction between rotor and stator and,
generally speaking, it could be divided as:

Figure 6.3 Flow Structures with 5 to 6 Orders of Magnitudes Variations in Length and Time
Scales (LaGraff et al., 2006)

120 E. M. Greitzer, “Thermoaldynamics and fluid mechanics of turbomachinery”, AS1/E 9713, NATO, 1985.
126

1. Interaction of potential flows in adjacent blade rows including Transient Fan.


2. Interaction between the wake flow and blade rows downstream.
3. Interaction between the secondary flows and blade rows.
4. Interaction wake-boundary layer.
5. Un-shrouded tip leakage flow interaction.
6. Film Cooling effects.

6.3 Interaction of Potential Flows in Adjacent Blade Rows


The first part comes from the changing of the relative position of rotor to stator which results in the
periodic fluctuation of the pressure or shocks. This fluctuation is propagated both upstream and
downstream as disturbance waves.

A - Mach
number
contours

B - Install C - Chocking D - Near Pick Effeciency

Figure 6.4 Shock Structure in Transonic Fan

6.3.1 Interactions in Transonic Fan


The shock structure associated with transonic fans is complicated by the 3D nature of the flow field
127

and operating range over which the fan must operate121. Figure 6.4 (A-B-C) illustrates some typical
features – leading edge oblique shock, aft passage normal shock below peak efficiency, and a near-
normal, detached bow shock near peak efficiency (and higher) loading conditions. Note that
throughout this report, loading refers to flow turning. For high tip-speed fans (inlet relative Mach
numbers greater than 1.4), the trend seems to be to design for an oblique leading edge shock through
higher loading conditions (near and at peak efficiency). This trend seems reasonable given the
continued need to reduce losses. Other flow field considerations in transonic fans include the
interrelationship between the rotor tip-clearance vortex structure and passage shock, high Mach
number stator flow, most notably in the hub region, and strong shock – boundary layer interaction.

6.4 Interaction Between Wake Flow and Blade Rows


The second part, unsteady wake, is a quite common flow phenomenon, not only in turbomachinery.
Due to the thickness of the trailing edge of blade, the flows after the blade generate a high dissipation
region, called wake, which is similar to the flow passed a circular cylinder where a famous wake flow
Von Karman Vortex Street can be observed. When a viscous flow passes a cylinder or an airfoil, a
regular vortex shedding can be found behind the
cylinder, which results in a zone with fully turbulent
flow and high dissipation. The pressure on the surface
of cylinder will fluctuate with the vortex shedding. A
similar flow phenomenon exists in the bypass flow
after a blade. Figure 6.5 (Wang and He, 2001), shows
the results of unsteady simulation performed by Wang
and He, in which the instantaneous pressure contour
patterns of wake for turbulent flow through unsteady
simulations are presented clearly. The wake flow in
multi-stage turbomachinery is more complicated than
vortex shedding after circle cylinder since it will be
distorted and deformed by the blade when flows
through the blade row downstream as shown clearly
in Figure 6.7 by [Smith, 1966; Stieger & Hodson,
2005]. This unsteady transport process could last to Figure 6.5 Pressure Contour of Wake
the next few blade rows and mix with new wake flows Flow
to forming highly non-uniform unsteady flow in blade passage.

6.5 Interaction Between Secondary Flows and Blade Rows


The third part is similar to the second one, in which the second flows are also sheared by the blade
rows downstream during the transport process. The distortion and mixing of these vortices will
enhance the non-uniformity of the flow. [Schlienger et al.] investigated the interaction between
secondary flows and blade rows through experiments on a low speed turbine with two stages. It is
found that the characteristic of the unsteady flow field at the rotor hub exit is primarily a result of
the interaction between the rotor indigenous passage vortex and the remnants of the secondary flow
structures that are shed from the first stator blade row. Moreover, there exist interactions among
secondary flows, wake and blade rows, which results in more complicated unsteady flow.
[Matsunuma]122 investigated this interaction effect on a low speed turbine of single stage, with the

121Boyer, K., M., “An Improved Streamline Curvature Approach for Off-Design Analysis of Transonic Compression
Systems “, PhD. Dissertation, Virginia Polytechnic Institute and State University, 2001.

T.Matsunuma, “Unsteady flow field of an axial-flow turbine rotor at a low Reynolds number”, ASME-GT06,
122

number 90013, Spain, 2006.


128

instantaneous absolute velocity contour pattern at the nozzle exit shown in Figure 6.6. The
experimental results suggest that the secondary vortices are periodically and three-dimensionally
distorted at the rotor inlet. A curious tangential high turbulence intensity region spread at the tip
side is observed at the front of the rotor, which is because of the axial stretch of the nozzle wake due
to the effects of the nozzle passage vortex and rotor potential flow field.

Figure 6.6 Instantaneous Absolute Velocity Contour at Nozzle Exit [Matsunuma, 2006]

6.6 Wake-Boundary Layer Interaction


In low-pressure turbines, the wakes from upstream blade rows provide the dominant source of
unsteadiness. Under low Reynolds number conditions, the boundary-layer transition and separation
play important roles in determining engine performance. An in-depth understanding of blade
boundary layer spatial-temporal evolution is crucial for the effective management and control of
boundary layer transition or separation, especially the open separation, which is a key technology
for the design of low-pressure turbines with low Reynolds number. Thus it is very important to
research the wake-boundary layer interaction. In low-pressure turbines with low Reynolds number,
boundary layer separation may occur as the blade load increases. Rational use of the upstream
periodic wakes can effectively inhibit the separation by inducing boundary layer transition before
laminar separation can occur, so as to control loss generation. A comprehensive and in-depth
research of wake boundary layer interactions in low-pressure turbines is given by [Hodson & Howell
(2005)]. They summarized the processes of wake-induced boundary-layer transition and loss
generation in low-pressure turbines. The periodic wake-boundary layer interaction process is as
follows (see Figure 6.7):
129

• When the wake passes, the wake-induced turbulent spots form within attached flows in front of
the separation point, the turbulent spots continue to grow and enter into the separation zone,
and consequently inhibit the formation of separation bubble. The calmed region trails behind
the turbulent spots. It is a laminar-like region, but it has a very full velocity profile. The flow of
the calmed region is unreceptive to disturbances. Consequently, it remains laminar for much
longer than the surrounding fluid and can resist transition and separation. It is the combination
of the calming effect and the more robust velocity profile within the calmed region that makes
this aspect of the flow so important. After the interaction of the wake, boundary layer separation
occurs in the interval between the two wakes.

Figure 6.7 Unsteady Wakes Convecting in Blade Passage

6.7 Un-Shrouded Tip Leakage Flow Interaction


The tip leakage flow is important in most turbomachinery, where a tip clearance with a height of
about 1-2% blade span exists between the stationary end wall and the rotating blades. An
unshrouded tip design is widely employed for a low stress and/or a better cooling in modern high-
pressure turbines. Pictorial representation of the tip leakage flow in unshrouded blades is given in
Figure 6.8 (left and right). The leakage flow over unshrouded blades occurs as a result of the
pressure difference between the pressure and suction surfaces and is dominated by the vortex shed
near the blade tip. The tip leakage flow has significant effects on turbomachinery in loss production,
aerodynamic efficiency, turbulence generation, heat protection, vibration and noise. As a
consequence of the viscous effects, significant losses are generated by the tip leakage flow in regions
inside and outside the tip gap. And the entropy creation is primarily due to the mixing processes that
take place between the leakage flow and the mainstream flow. [Denton (1993)] gave a simple
prediction model for the tip leakage loss of unshrouded blades. So far, there are many researches
130

about the leakage flow unsteady interactions in compressor. For example, [Sirakov & Tan (2003)]
investigated the effect of upstream unsteady wakes on compressor rotor tip leakage flow. It was
found that strong interaction between upstream wake and rotor tip leakage vortex could lead to a
performance benefit in the rotor tip region during the whole operability range of interest. The
experimental result of [Mailach et al. (2008)] revealed a strong periodical interaction of the incoming

Figure 6.8 Flow Over an Unshrouded Tip Gap

stator wakes and the compressor rotor blade tip clearance vortices. As a result of the wake influence,
the tip clearance vortices are separated into different segments with higher and lower velocities and
flow turning or subsequent counter-rotating vortex pairs. The rotor performance in the tip region
periodically varies in time. Compared with in compressor, very little published literature is available
on the unsteady interactions between leakage flows and adjacent blade rows in turbine. [Behr et al.
(2006)] indicated that the pressure field of the second stator has an influence on the development of
the tip leakage vortex of the rotor. The vortex shows variation in size and relative position when it
stretches around the stator leading edge.

6.8 General Review on Secondary Flows


The important 3D viscous flow phenomena within a blade passage of turbomachinery are boundary
layers and their separations, tip clearance flows and wakes, which are most responsible of energy
losses existing in blade passage. Hence, the losses in an axial compressor or turbine can be mainly
classified as123:

• Profile losses due to blade boundary layers and their separations and wake mixing; in high
speed condition, shock/boundary layer interaction may exist.
• End-wall boundary layer losses, including secondary flow losses and tip clearance losses.
• Mixing losses due to the mixing of various secondary flows, such as the passage vortex and
tip leakage vortex.

Among all these losses, the most complex one is the secondary flow loss. That is why considerable
research on the secondary flow phenomena has been done in last decades. Secondary flow is
defined as the difference between the real flow and a primary flow, which is related to the
development of boundary layer on end-wall and blade surface, the evolution of vortices in
passage, and detached flows or simply, the secondary flow in a blade row can be defined as any

123 Sh. Kang, “Investigation on the Three Dimensional within a Compressor Cascade with and without Tip
Clearance”, PhD thesis, Vrije Universiteit Brussel, September 1993.
131

flow, which is not in the direction of the primary or stream wise flow 124. Based on topology
analysis and experiments, as well as the numerical simulations in recent decades, a couple of
secondary flow models are proposed which are presented below.
6.8.1 Classical View
The so-called classical secondary flow model, as illuminated in Figure 6.9 (a-b), is proposed by

(a) Classical View


(Hawthorne, 1955)

(b) Secoundary Losses in


presence of secoundary
vortex flow in classical
view

Figure 6.9 Classical Secondary Flow Model

124 Lei Qi and Zhengping Zou, “Unsteady Flows in Turbines”, Beihang University China.
132

Hawthorne125 for the first time according to the theory of inviscid flow in 1955. This model presents
the components of vorticity in the flow direction when a flow with inlet vorticity is deflected through
a cascade. The main vortex, so-called passage vortex, represents the distribution of secondary
circulation, which occurs due to the distortion of the vortex filaments of the inlet boundary layer
passing with the flow through a curved surface. The vortex sheet at the trailing edge is composed of
the trailing filament vortices and the trailing shed vorticity whose sense of rotation is opposite to
that of the passage vortex. The classical vortex model attributes the secondary flow losses to the
generation and evolution of vortex passage. However, this model is relatively simple, in which the
interaction between the inlet boundary layer and blade force was not considered. Moreover, the
vortex system within passage is only single passage vortex in half of the passage height range with
other vortices absences. The secondary flow losses can be visualized by absence/presence of
secondary vortex on
Figure 6.9 (b).
6.8.2 Modern View
When a shear flow along
the solid wall approaches a
blade standing on the wall,
the shear flow will be
separated from the wall
and roll up into a vortex in
front of the blade leading
edge. This vortex is called
horseshoe vortex due to its
particular shape. This well-
known phenomenon is
firstly observed in the flow
around cylinders. The oil
(a) Kline 1966
flow visualizations by
[Fritsche]126 show the
evidence of the horseshoe
vortex in accelerating
cascades. In 1966, Klein
presents a finer cascade
vortex model with both the
passage and horseshoe
vortices as depicted in
Figure 6.10 (a). While, the
pioneering work for
detailed analysis of
secondary flow patterns in
turbine cascades in general (b) Langston, 1977
is done in 1977 by
[Langston et al.]127 who Figure 6.10 Modern Secondary Flow Model
proposed the well-known

125 W. R. Hawthorne,” Rotational flow through cascades part 1: the components of vorticity.” Journal of Mechanics
and Applied Mathematics, 8(3):266–279, 1955.
126 A. Fritsche. Str¨omungsvorg¨ange in schaufelgittern. Technische Rundschau Sulzer, 37(3), 1955.
127 L. S. Langston, “Three-dimensional flow within a turbine blade passage”, Journal of Engineering for Power,

99(1):21–28, 1977.
133

modern vortex model in


cascade. Three vortices
are presented in this
model, as depicted in
Figure 6.11 (b).
Langston explains the
interaction between the
horseshoe vortex and
the passage vortex, and
the development of the
passage vortex. The big
differences between
Langston’s model and
Klein’s model exist in
twofold128: by Langston (a) Sharma and Butler, 1987
et al129 who proposed the
well-known modern
vortex model in cascade.
Three vortices are
presented in this model,
as depicted in Figure
6.11 (b). Langston
explains the interaction
between the horseshoe
vortex
and the passage vortex,
and the development of
the passage vortex. The
big differences between
Langston’s model and
Klein’s model exist in (b) Goldstein and Spores, 1988
twofold130:
Figure 6.11 Vortex pattern of Latest Secondary Flows
• Langston clearly
postulates that the pressure side leg of the leading edge horseshoe vortex, which has the same
sense of rotation as the passage vortex, merges with and becomes part of the passage vortex
• Langston clarifies that the suction side leg of the leading edge horseshoe vortex which rotates
in the opposite sense to the passage vortex, continuing in the suction side end-wall corner,
while the presentation of Klein suggests that this vortex is gradually dissipated in contact
with the passage vortex.

The first point from Langston is supported by the light sheet experiment by Marchal and

128 C. H. Sieverding, “Recent progress in the understanding of basic aspects of secondary flows in turbine blade
passages”, Journal of Engineering for Gas Turbines and Power, 107(2):248–257, 1985.
129 L. S. Langston, “Three-dimensional flow within a turbine blade passage”, Journal of Engineering for Power,

99(1):21–28, 1977.
130 C. H. Sieverding, “Recent progress in the understanding of basic aspects of secondary flows in turbine blade

passages”, Journal of Engineering for Gas Turbines and Power, 107(2):248–257, 1985.
134

Sieverding131 in 1977. While, the results of this experiment also show the counter-rotating vortex,
called counter vortex by Langston, in the trailing edge plane on the mid span side of the passage
vortex rather than in the corner, which is not consistent with the second point from Langston.
6.8.3 Latest View
In 1987, [Sharma and Butler]132 proposed a secondary flow pattern which is slightly different to that
from Langston. This pattern, shown in Figure 6.11 (a), demonstrates that the suction side leg of the
horseshoe vortex wraps itself around the passage vortex instead of adhering to the suction side. This
result is similar to the results of [Moore]133 and [Sieverding]134. However, in 1988, another pattern is
given by [Goldstein and Spores]135, shown in Figure 6.11 (b), which is different to Sharma’s again.
Based on mass transfer results, they suggested that the suction side leg of the horseshoe vortex stays
above the passage vortex and travels with it. This flow pattern is similar to that suggested by [Jilek]136
in 1986. The major difference among these three models is the location of the suction side leg of the
horseshoe vortex. Since it is difficult to be detected due to the small size, most literatures cannot
demonstrate develop of this vortex clearly. In 1997, a very detailed secondary flow visualization
study was performed by Wang137. They proposed a more comprehensive but more complicated
secondary flow the passage vortex and travels with it. This flow pattern is similar to that suggested
by Jilek138 in 1986. The major difference among these three models is the location of the suction side
leg of the horseshoe vortex. Since it is difficult to be detected due to the small size, most literatures
cannot demonstrate develop of this vortex clearly. In 1997, a very detailed secondary flow
visualization study was performed by [Wang]139. They proposed a more comprehensive but more
complicated secondary flow pattern, as illustrated in which includes the passage vortex, the
horseshoe vortex, the wall vortex and the corner vortex. The development of the horseshoe vortex
nearby the end-wall is effected by the boundary layer on end wall and the blade surface. In modern
advanced blade, the leading edge radius of blade is so small that can be compared with the thick of
boundary layer. Hence, the separation of boundary layer on end wall generates the multi-vortex
structures at the leading edge of blade. Due to a strong pressure gradient the pressure side leg of the
horseshoe vortex moves toward the suction side after it enters the passage. Meanwhile it entrains
the main flow and the inlet boundary layer forming a multi-vortex leg. In 2001, [Langston]140
reviewed these new models after the [Sieverding’s] review. Laster in the same year, [Zhou and
Han]141 gave a more comprehensive review of all these models. They concluded that the good

131 P. Marchal and C. H. Sieerding, “Secondary flows within turbomachinery blading’s”, CP 214, AGARD, 1977.
132 O. P. Sharma and T. L. Butler, “Prediction of the end wall losses and secondary flows in axial flow turbine
cascade. Journal of Turbomachinery”, 109:229–236, 1987.
133 J. Moore and A. Ransmayr, “Flow in a turbine cascade part 1: losses and leading edge effects”, ASME, 1983.
134 C. H. Sieverding and P. Van den Bosch,” The use of colored smoke to visualize secondary flows in a turbine-

blade cascade”, Journal of Fluid Mechanics, 134:85–89, 1983.


135 R. J. Goldstein and R. A. Spores, “Turbulent transport on the end wall in the region between adjacent turbine

blades”, Journal of Heat Transfer, 110:862–869, 1988.


136 J. Jilek, “An experimental investigation of the three-dimensional flow within large scale turbine cascades”,

ASME-GT86, number 170, 1986.


137 H. P. Wang, S. J. Olson, R. J. Goldstein, and E. R. G. Eckert, “Flow visualization in a linear turbine cascade of high

performance turbine blades”, Journal of Turbomachinery, 119(1):1–8, 1997.


138 J. Jilek, “An experimental investigation of the three-dimensional flow within large scale turbine cascades”,

ASME-GT86, number 170, 1986.


139 H. P. Wang, S. J. Olson, R. J. Goldstein, and E. R. G. Eckert, “ Flow visualization in a linear turbine cascade of

high performance turbine blades”, Journal of Turbomachinery, 119(1):1–8, 1997.


140 L. S. Langston, “Secondary flows in axial turbines: a review”, Annals of the New York Academy of Sciences, 934

(Heat Transfer in Gas Turbine System):11–26, 2001.


141 X. Zhou and W. J. Han, “A review of vortex model development for rectangular turbine cascade”, (in Chinese).

Journal of Aerospace Power, 16(3):198–204, 2001.


135

understanding of the secondary flow in turbomachinery can help greatly to control the vortices
within passage and decrease the losses, help greatly to control the vortices within passage and
decrease the losses.
6.8.4 Comparing and Contrasting Secondry Flow in Turbine and Compressors
Another view begins by comparing and contrasting turbine and compressor secondary flows,
together with conclusions on the way forward to design in compressors142. A large amount of
material has been published on secondary flow effects in axial flow turbomachinery, both turbines
and compressors. Only a brief summary of these is given here. As will be seen in the next section,
non-axisymmetric end wall profiling has been pursued in recent years principally in the field of axial
flow turbines. Consequently, it is useful to compare and contrast turbine and compressor secondary
flows.
Comprehensive reviews of
turbine secondary flows are
given in [Sieverding and
Langston], and of secondary
loss generation in [Denton].
Whilst secondary flows are
induced by any total
pressure profile that enters a
blade row and is
subsequently deflected by it,
the clearest understanding
has been obtained for the
case when the total pressure
profile is just due to the
incoming end wall boundary
layers. Figure 6.12 shows a
diagrammatic
representation of turbine
end wall secondary flows
taken from [Takeishi et al.]
(note that the rotation of the
vortices is generally
exaggerated) which has
been describe this more Figure 6.12 Turbine Secondary Flow Model (Takeishi et al.)
fully, but the basic elements
are:
• Rolling up of the inlet boundary layer into the horseshoe vortex at the airfoil leading edge.
The pressure surface side leg of this becomes the core of the passage vortex. The passage
vortex is the dominant part of the secondary flow and beneath it on the end wall a new
boundary layer is formed, referred to as cross-flow "B" in Figure 6.12, which starts in the
pressure side end wall corner.
• Upstream of this the inlet boundary layer is deflected across the passage (over turned),
referred to as cross-flow "A". The end wall separation line marks the furthest penetration of
the bottom of the inlet boundary layer into the passage and divides it from the new boundary
layer forming downstream of it. The dividing streamline between the suction and pressure

142N W Harvey, “Some Effects of Non-Axisymmetric End Wall Profiling on Axial Flow Compressor Aerodynamics.
Part I: Linear Cascade Investigation”, Proceedings of GT2008.
136

side flows is shown as the attachment line in Figure 6.12. It intersects with the separation
line at the saddle point.
• The new end wall boundary layer, cross-flow "B", carries up onto the airfoil suction surface
until it separates (along the airfoil "separation line") and feeds into the passage vortex. The
suction side leg of the horseshoe vortex, referred to as the counter vortex in Figure 6.12,
remains above the passage vortex and moves away from the end wall as the passage vortex
grows.
• A small corner vortex may occur in the suction surface/ end wall corner rotating in the
opposite sense to the passage vortex. This has the effect of opposing the overturning at the
end wall, although at the cost of additional loss.
One additional source of “classical” secondary flow that must be mentioned is the trailing edge
vorticity that originates as a vortex sheet downstream of the blade trailing edge due to the variation
in circulation along the span of the airfoil (and not shown). The scope for reducing this by modifying
the end wall flows does not appear to be great and has not been part of this study. The basic features
of compressor secondary flows are the same as those in a turbine blade row. However, there are a
number of important differences in the details between the two, [Cumpsty]143:
• The turning in a compressor blade row is much lower; typically 30 – 40 degree , compared to
100 degree in a turbine.
• From classical secondary flow theory, this would be expected to result in lower secondary
flows in a typical compressor row, for a comparable inlet total pressure profile.
• An additional feature, often overlooked, for turbine secondary flows is that once they have
rolled up into vortices any further acceleration of the flow will stretch them feeding in more
kinetic energy (of rotation), [Patterson]. This may have the effect of amplifying the benefit of
anything that delays the initial development of secondary flows on the end walls.
• Since the flow through a compressor blade row diffuses such vortex stretching will not occur.
Rather the diffusion will encourage more rapid mixing out of the vortices. It is suggested that
this is the reason why the smaller vortices (counter and corner) seen in turbine rows are not
often identified for compressor ones. In addition end wall over-turning in a compressor row
will be much more likely to result in flow separation, especially when the static pressure rise
across the row increases if the compressor moves up its characteristic.

6.9 3D Separation
A number of different flow regimes come under the heading of “three-dimensional separation”:
• If the aerodynamic loading is low enough, then the low momentum fluid in the airfoil suction
side/ end wall corner will separate off the blade surfaces (as in turbine secondary flows) but
will still have forward momentum.
• Where the loading is such that reverse flow does occur, then this may initially only be on one
of either the end wall or the airfoil suction surface refers to the former as “wall stall” and the
latter as “blade stall”.
• The combination of these two is known as “corner stall”. The resulting flow patterns are
illustrated in Figure 6.13. where the illustration of formation of hub corner stall together
with limiting streamlines and separation lines, (Lei et al.).

143 Cumpsty N. A.,, (2004), “Compressor Aerodynamics”, Krieger Publishing Company.


137

Distinct features of this are the


reverse flow on both walls and
the decrease of the chord wise
extent of this flow away from
the end wall. In terms of
secondary loss, it is difficult to
generalize on its magnitude in
compressor rows. This depends
on the details of the design; of
which diffusion factors,
DeHaller numbers and aspect
ratio are just a few. One example
may serve to indicate the
potential for losing
aerodynamic performance.
With a small leakage flow
present, which suppressed the
corner stall, the 54% was
reduced to 13% (about 11% of
the total). For a turbine row
with a similar aspect ratio, the
secondary losses may be Figure 6.13 Illustration of formation of hub corner stall together
expected to be at least 20% of with limiting streamlines and separation lines
the total, but again this depends
on the design details. From the
above it is concluded that the scope
for reducing secondary loss in a well-
designed compressor row at its
design condition (without corner
stall) is likely to be less than for a
typical turbine one. Rather, reducing
or mitigating penalizing features such
as corner stall may be of more
importance to the compressor
aerodynamic designer.
6.9.1 Compressors Example
As an example, consider the small
variations in leading edge geometry,
leading edge roughness, leading
edge fillet, and blade fillet geometry
on the three-dimensional separations
found in compressor blade rows, as
investigated by [Goodhand and
Miller]144. The detrimental effects of
these separations have historically Figure 6.14 Three-dimensional separations: traditional
been predicted by correlations based view and scope of current investigation

144 Martin N. Goodhand and Robert J. Miller, “The Impact of Real Geometries on Three-Dimensional Separations
in Compressors”, Journal of Turbomachinery, 2012.
138

on global flow parameters, such as blade loading, inlet boundary layer skew, etc., and thus ignoring
small deviations such as those highlighted above. The results show that any deviation which causes
suction surface transition to move to the leading edge over the first 30% of span will cause a large
growth in the size of the hub separation, doubling its impact on loss. The geometry deviations that
caused this, and are thus of greatest concern to a designer, are changes in leading edge quality and
roughness around the leading edge, which are characteristic of an eroded blade. 3D separations
always occur on compressor blades in the corner between the suction surface and the end wall145
(see Figure 6.14). Blades can be designed such that these are relatively small and benign; however,
as loading or incidence is increased, their size and thus detrimental effect can increase significantly.
In practice, it is these separations which limit the total blade loading by their impact on loss, blockage,
and deviation.

6.10 Airfoil End-Wall Heat Transfer


The flow in a gas turbine influenced by the inner hub and outer casings of the airfoils is defined as
secondary or end-wall flows146. These flows often contain vortices that give rise to velocity
components that are orthogonal to the primary flow direction as depicted in Figure 6.15 by the
ribbon arrows, which is specific to vane end-walls147. These flows constitute one of the most common
place and wide spread three-dimensional flows arising in axial flow turbomachinery. In typical
modern day turbine designs, end-wall flows for first stage vanes are responsible for over 30% of the
total pressure loss through a turbine stage leading to a reduction in turbine efficiencies on the order
of 3%. While overall airfoil losses
have been reduced through the use
of three dimensional geometries
that make use of bowed or leaned
airfoils, the end-walls have
remained fairly conventional and
the source of much of the
remaining pressure losses.
The heat transfer consequences
are immense because of the
increased convective coefficients
and mixing out of film-coolant near
the surface. It is clear from a
thermodynamic analysis of a
turbine engine, that to improve
performance, there is a need to
increase the aspect ratio for
turbine airfoils and to increase
turbine inlet temperatures. To
improve a turbine’s performance,
these trends require that end-wall Figure 6.15 Illustration of the near wall flows as taken through
flows be carefully considered in oil and dye surface flow visualization (reproduced with
turbine designs. The end-wall flow permission of the publisher from ASME)
through an airfoil cascade under
isothermal conditions with an approaching two-dimensional boundary layer agrees well with that

145 Gbadebo, S. A., Cumpsty, N. A., and Hynes, T. P., 2005, “Three-Dimensional Separations in Axial
Compressors,” ASME J. Turbomachinery , 127, pp. 331–339.
146 Karen Thole, “Mechanical Engineering Department Penn State University”, Park, PA 16802-1412.
147 Langston, L. S. “Cross flows in a Turbine Cascade Passage,” ASME J of Engineering for Power 102 -1980.
139

depicted in Figure 6.16-(Top). The flow model shows that the inlet boundary layer separates from
the approaching end-wall to form what is known as a horseshoe vortex. One leg of the horseshoe
vortex, present on the pressure side of the airfoil (concave side), is convected into the passage and is
promoted by the inherent pressure gradient between the two airfoil surfaces. This pressure side leg
of the horseshoe vortex develops into what is known as the passage vortex. The other leg of the
horseshoe vortex, present on the suction side of the airfoil (convex side), has an opposite sense of
rotation to the larger passage vortex and develops into what is known as a counter vortex. The
counter vortex can be thought of as a planet rotating about the axis of the passage vortex (sun). While
this picture represents a time averaged representation, measured data indicates that the vortex is
not steady. While the development of the vortical structures originates in the end-wall regions, the
growth can be such that the passage vortex occupies a large portion of the airfoil exit. This vortical
structure growth extends up to 30-40% of the total span for older vane designs and has been reduced
to approximately 10-15% of the total span
in the last 15 years. The flow patterns
previously described make it difficult to
cool the end-wall, particularly when
considering the near end-wall flow, as
illustrated in Figure 6.16148. The surface
flow visualization in Figure 6.16, achieved
through an oil and surface dye technique,
illustrates the strong cross flows that occur.
The cross flows are driven by the inherent
pressure gradients from the pressure to the
suction side of adjacent airfoils. For
example, these cross flows influence how
the film-cooling jets exit from the holes as
well as influence the end-wall heat transfer
coefficients149.
6.10.1 Theoretical Development of End-
Wall Flows
As the end-wall boundary layer approaches
a turbine airfoil, the flow stagnates,
whereby the total pressure becomes the
static pressure along the span of the vane.
Given that the fluid nearer to the end-wall
has a lower velocity, a stronger deceleration
in the boundary layer occurs for the higher
speed fluid than for the lower speed fluid. As
a result of these differences in the
deceleration, a transverse static pressure
gradient occurs along the vane span causing
Figure 6.16 (Top) - Measurements of the Horseshoe
the higher speed fluid to turn toward the Vortex just upstream of the Vane at the Vane-End-Wall
end-wall plate. Subsequently, the formation Juncture (Bottom) - Actual Hardware Showing Effects
of a horseshoe vortex occurs just upstream of the Horseshoe Vortex on a First Vane (Courtesy of
ASME)

148 Friedrichs, S., Hodson, H. P. and Dawes, W. N., “Distribution of Film-Cooling Effectiveness on a Turbine End-
wall Measured Using the Ammonia and Diazo Technique,” J of Turbomachinery 118 -1996.
149 Karen Thole, “Mechanical Engineering Department Penn State University”, Park, PA 16802-1412.
140

of the turbine vane. One of the legs of the horseshoe vortex wraps around the pressure side of the
vane and the other leg wraps around the suction side of the vane. Figure 6.16 (top) shows
measurements of the horseshoe vortex upstream
of the turbine vane and Figure 6.16
(Bottom)150 shows extreme damage from the
effects of this vortex on actual hardware151.
These measurements were made for a simple
case with an isothermal flow and an approaching
boundary layer that was 9 % of the vane span
(Z/S = 0.09). From the combined contour/vector
plot in Figure 6.16 (top), one can see where the
flow separates from the upstream end-wall and
how the flow is then rolled into a vortex. It is also
clear to see that, if a film-cooling jet were injected
into this region, it would be difficult to maintain
the coolant along the end-wall.
The flow fields discussed thus far have described
the condition for a uniform, isothermal flow field
with an approaching two-dimensional boundary
layer along the end-wall. In practice, this
idealized flow situation rarely happens since an
upstream combustor is present whereby there
can be large variations in the exiting flow. Non-
uniformity of inlet profiles in addition to the
viscous boundary layers along the end-wall are
caused by temperature gradients at the
combustor exit. The development of crossflow
and vortical motions in a curved passage, such as
an airfoil passage, can be understood by
considering flow along two streamlines, as
shown in Figure 6.17. Two idealized cases are
considered: Top) a gradient of velocity due to a
turbulent inlet boundary layer with an Figure 6.17 Illustration of Different Vortical
isothermal flow, Bottom) a linear temperature Patterns that are possible for two idealized flow
gradient typical at the exit of a combustor with a conditions: Top) Isothermal with an Inlet
uniform velocity field152. Assuming steady, Boundary Layer ; and , Bottom) Inviscid Flow
incompressible, inviscid flow with negligible with a Temperature Profile
variation of velocity in the n-direction, the
centripetal acceleration for the streamlines A and B must be balanced by the pressure gradient across
the pitch:
2 2
∂p VSA ∂p VSB
| =ρ , | =ρ
∂n A RA ∂n B RB

150 Friedrichs, S., Hodson, H. P. and Dawes, W. N., “Distribution of Film-Cooling Effectiveness on a Turbine End-
wall Measured Using the Ammonia and Diazo Technique,” J of Turbomachinery,1996.
151 M. Kang, A. Kohli, and K.A. Thole, “Heat Transfer and Flow-field Measurements in the Leading Edge Region of

a Stator Vane End-wall,” J of Turbomachinery,1999.


152 B. Lakshminarayana, “Fluid Dynamics and Heat Transfer of Turbomachinery”, (NY: Wiley Inter-science,

1996).
141

Eq. 6.1
From boundary layer theory ∂p/∂n, for both streamlines from the developing boundary layer on the
sidewalls should be equal. Therefore
∂p ∂p
| = |
∂n A ∂n B
Eq. 6.2
Due to the viscous turbulent boundary layer at the end-wall, it is evident that VsB < VsA. So for the
boundary layer assumption to hold, the radius of curvature of the streamline at B must be reduced.
This creates a crossflow in the boundary layer from the pressure surface towards the suction surface
of the blade, and thus generates secondary flow (flow that is not aligned with the stream-wise
direction) as depicted in Figure 6.17(top). Now consider Figure 6.17(Bottom), a constant
velocity profile with linear temperature profile. The same physics hold for this case; however, the
resulting vortex is reversed in direction. In this instance the temperature at A is greater than at B,
therefore ρA < ρB. Now RB must be greater than RA for the normal pressure gradient to balance and
the cross flow is generated towards the pressure side of the adjacent vane row. The change in
streamline curvature would be less severe in this case compared to Figure 6.17(top) since the
velocity term is squared in the relationship of Eq. 6.1.
As one can see from these simple, idealized flow situations, there can be large variation in the
expected secondary flow pattern that can be derived in a turbine vane passage. The important driver
for how the flow develops in a turbine vane passage is the total pressure profile entering the passage.
As this total pressure profile becomes the static pressure along the vane stagnation, the flow will be
driven from a high pressure region to a low
pressure region. In most turbine designs there is
a flow leakage slot between the combustor and
the turbine whereby cooler fluid is injected into
the main hot gas path. This leakage can also have
an effect on the secondary flow patterns that
develop. The bottom line when considering end-
wall flows is that the profile exiting the
combustor, which is often referred to as the
combustor pattern factor that ultimately enters
the turbine, should be known to fully predict the
secondary flows that will develop in the turbine
passage. In practice, the combustor pattern
factor is one of the parameters used in designing
cooling schemes for the airfoils and their
associated platforms.
6.10.2 End-Wall Heat Transfer
The heat transfer coefficients given in Figure
6.18 are represented in terms of a non-
dimensional Stanton number based on exit
velocity. In the region upstream of the vanes,
there is a high heat transfer region that occurs Figure 6.18 Contours of Non-Dimensional Heat
Transfer Coefficients (Reproduced with
between the stagnation point and the
Permission ASME)
reattachment of the flow on the suction side of
the airfoil. This is the area which experiences
very high acceleration. As the flow moves through the passage, it is apparent that the location of the
peak Stanton numbers (peak heat transfer) is being swept from the outer pressure surface towards
142

the suction side of the central vane153. Note that the heat transfer data shown here were taken with
a two-dimensional inlet boundary layer under low speed conditions, to allow for highly-resolved
data, with matched Reynolds number conditions. There is evidence in the literature that the
secondary flow patterns remain the same at both low and high Mach number conditions. Rather,
these secondary flows are a stronger function of the airfoil geometry and inlet profile conditions that
of Mach number. There is no data in the literature that discusses how the heat transfer on the end-
wall is altered depending upon an inlet flow condition that is relevant to that exiting a combustor. As
was stated previously, it is important to consider that the profile exiting the combustor can vary
greatly from that of a two-dimensional boundary layer assumption. For further info and cooling
effects, please see chapter 7.
6.10.3 Leading Edge Modifications
Because industry is concerned with problems at the vane leading edge-end-wall juncture, a number
of more recent studies have begun to evaluate geometric modifications to airfoils in this region. At
this point there have been three different geometric concepts tested for an asymmetric airfoil
geometry that include the following classifications: fences, fillets, and bulbs. Methods using flow
control such as suction combined with injection have also been reported but are generally not
feasible for gas turbines where gas temperatures exceed airfoil melting temperatures. [Chung and
Simon] first presented their concept for secondary flow control in 1993 that encompassed using a
fence placed in mid-passage between two turbine airfoils154. While their tests indicated a reduction
in strength of the passage vortex, industry’s concern was in cooling the fence and that it acted as a fin
conducting heat to the platform as it was exposed to the hotter main gas path fluid.
Three-dimensional end-wall contouring, which includes a more comprehensive geometric
modification than simply a modification to the end-wall-airfoil leading edge juncture, has also been
investigated computationally by Harvey, et al. and experimentally verified by [Hartland, et al.]155. To
design the end-wall contour, they used a linear
sensitivity matrix in conjunction with superposition
methods prior to applying an inverse design
algorithm. The results of the experimental
verification confirmed a predicted reduction in exit
flow angle deviations. Moreover, the experiments
indicated a 30% reduction in loss, which was higher
than predicted. In a later study, [Brennan, et al. and
Rose, et al.] applied similar computational and
experimental (respectively) methodologies as
[Harvey et al. and Hartland, et al.]. They applied
these methods to a high pressure turbine for a single
stage in both the vane and blade passages. They Figure 6.19 Fillet and Bulb Designs as Shown
reported stage efficiency improvements of 0.59%, by (Becz et al.)
which exceeded their predicted improvement of
0.4%. Using end-wall contouring and leading edge modifications show promise in reducing
secondary flows; however, there are numerous effects that need to be considered. Because this

153 M. Kang, A. Kohli, and K.A. Thole, “Heat Transfer and Flow field Measurements in the Leading Edge Region of
a Stator Vane End-wall,” J of Turbomachinery, 1999.
154 J. T. Chung and T. W. Simon, “Effectiveness of the Gas Turbine End-wall Fences in Secondary Flow Control at

Elevated Freestream Turbulence Levels,” ASME ,1993.


155 J. C. Hartland, P. G. Gregory-Smith, N. W. Harvey, and M. G. Rose, “Non-axisymmetric Turbine End Wall Design:

Part II – Experimental Validation,” J of Turbomachinery 122 (2000), ;Neil W. Harvey, Martin G. Rose, Mark D.
Taylor, Jonathan Shahrokh and David G. Groegory-Smith, “Non-axisymmetric Turbine End Wall Design: Part I –
Three-Dimensional Linear Design System,” J of Turbomachinery, 2000.
143

modification must be practically feasible, required manufacturing, space limitations, and cooling are
all practical issues that must be addressed. Moreover, some of these designs may be sensitive to the
inlet flow conditions which need to be considered156. See Figure 6.19.
6.10.4 Blade Tip Heat Transfer
The performance of a turbine engine is a strong function of the maximum gas temperature at the
rotor inlet. Turbine blade designers concentrate on finding adequate cooling schemes for high
pressure turbine blades, particularly the tip region where heat transfer is quite high. The clearance
between a blade tip and its associated outer casing, also known as the blade outer air seal, provides
a flow path across the tip that leads to aerodynamic losses and high heat transfer rates along the
blade tip. The flow within this clearance gap is driven by a pressure differential between the pressure
and suction side of the blade, and is also affected by the viscous forces as the fluid comes into contact
with the walls of the gap. A complete description of blade tips and the associated problems is given
by Bunker and is well described by Figure 6.20. Gap size, rotational effects, blade geometries and
Reynolds numbers were all highly influence the heat transfer coefficients.
One method for improving the thermal
environment along the blade tip is to inject
coolant into the tip region. In a review
paper on tip heat transfer, Bunker states
that for blade tips there have been very
few film-cooling studies reported in the
literature even though film-cooling is
widely used18. The discussion given in this
section is relevant to tip film-cooling (see
chapter 7) since that is what is typically
used in industry. Many tip heat transfer
and film-cooling studies have been
completed without rotational effects. In
general, there is evidence in the literature
that supports a widely variable effect of
rotation, which warrants further studies.
One of the first pioneers in this region is
[D. Metzger] where by [Kim et al.] presents
a summary of his work157. In addition to Figure 6.20 CFD Prediction of Streamlines Across a
concluding that there is only a weak effect Blade Tip (Reproduced With Permission From ASME).
of the relative motion between a
simulated blade and shroud on tip heat transfer coefficient, they stated that there is a strong
dependency of cooling effectiveness for a tip on the shape of the hole and injection locations. Four
hole configurations were discussed by [Kim et al.] that included the following: discrete slots located
along the blade tip, round holes located along the blade tip, angled slots positioned along the pressure
side and round holes located within the cavity of a squealer tip. The studies reported by [Kim et al.]
were performed in a channel that simulated a tip gap, whereby no blade with its associated flow field
was simulated. In comparing the discrete slots to the holes, their data indicated a substantial increase
in cooling effectiveness using the discrete slots for all blowing ratios tested. Injection from the
pressure side holes provided cooling levels of similar magnitude to the holes placed on the tip with

156Karen Thole, “Mechanical Engineering Department Penn State University”, Park, PA 16802-1412.
157Y. W. Kim and D. E. Metzger, “Heat Transfer and Effectiveness on Film Cooled Turbine Blade Tip Models,” J of
Turbomachinery 117 (1995);Y. W. Kim, J. P. Downs, F. O. Soechting, W. Abdel-Messeh, G. Steuber, and S.
Tanrikut, “A Summary of the Cooled Turbine Blade Tip Heat Transfer and Film Effectiveness Investigations
Performed by Dr. D. E. Metzger,” J of Turbomachinery 117-1995.
144

better spreading occurring in the case of the pressure side injection. Kim et al. also reported that an
increase in coolant mass flow generally yielded improved cooling with tip surface holes, but for
pressure side holes, increased coolant flow yielded decreased cooling effectiveness158.
6.10.5 Case Study 1 - Effects of Grid Refinement and Turbulence in 3D Flow Structure and End-
Wall Heat Transfer in Transonic Turbine Blade Cascade
Results of numerical simulation of 3D turbulent flow and end-wall heat transfer in a transonic turbine
cascade are presented by [Levchenya & Smirnov]159. Employing several turbulence models (k-ω
model by Wilcox, Menter SST model, v2-f model by Durbin), an analysis CFD predictability was done in
comparison with measurements in a linear cascade at the NASA Glenn Research Center transonic
turbine blade cascade facility. It has been concluded in particular that rather fine computational grids
are needed to get grid-
independent data on the end-
wall local heat transfer
controlled by complex 3D
structure of secondary flows.
With CFD codes of second-
order accuracy, one should
use grids comprised of about
or more than 2 M cells (for
each full blade passage) to get
a definite conclusion on
preference of one or another
turbulence model for
predictions of phenomena
under consideration. Figure 6.21 Blade Passage and Slice of The Computational Domain

6.10.5.1 Problem Definition


The geometry of the linear cascade is that available from the NASA GRC CD-ROM database arranged
by [Giel and Gaugler]160. A fragment of the cascade is illustrated in the Figure 6.21, together with a
slice of the computational domain. At the present computations, the fluid (air) is treated as a perfect
gas with the specific heat ratio γ = 1.4. The governing equations are the Reynolds-averaged Navier-
Stokes equations and the energy equation written for the total enthalpy. A power-law is adopted to
account for the dependency of viscosity on temperature, μ ∼ T 0.76. In order to define proper boundary
conditions at the 3D computational domain inlet section placed one axial chord upstream of the blade
leading edge, the (2D) turbulent flow developing in a parallel-plate channel was computed first,
assuming the adiabatic wall conditions. In the 2D flow computed separately for each of the turbulence
model used, a section was chosen that corresponded to the boundary layer thickness of 3.2 cm. Flow
field data at this section were used to define the total temperature, total pressure, velocity vector
angle and turbulence parameters distributions over the inlet plane of the 3D blade cascade
computational domain. To get the isentropic Mach number required a proper value of static pressure
was specified at the outlet boundary located one axial chord downstream of the blade trailing edge.
At the solid surfaces of the cascade the no-slip condition was imposed. The constant temperature, Tw,
of 350 K (that would correspond to power of 1560 Watts in the experimental prototype) was

158 Karen Thole, “Mechanical Engineering Department Penn State University”, Park, PA 16802-1412.
159 Alexander M. Levchenya and Evgueni M. Smirnov, “CFD-Analysis of 3d Flow Structure and End-wall Heat
Transfer in a Transonic Turbine Blade Cascade: Effects of Grid Refinement”, West-East High Speed Flow Field
Conference 19-22, November 2007.
160 Giel, P.W., and Gaugler, R.E., “NASA Blade 1. End-wall heat transfer data. Version 1,” NASA-Glenn Research

Center, Turbine Branch, CD ROM (2001).


145

specified on the end-wall, starting from the position of 0.3xCx (Cx referred to axial chord of blade)
upstream of the blade leading edge. Remaining walls were treated as adiabatic. Periodic boundary
conditions were used in the pitch wise direction. For computational purposes, only half of the real
span was considered, with the symmetry boundary condition at mid-span.
6.10.5.2 Computational Aspects
The 3D in house, incompressible/compressible Navier-Stokes code is based on the second order
finite-volume spatial discretization using the cell-centered variable arrangement and body-fitted
block-structured grids. For transonic flow analysis, a high-order version of the Jameson’s H-CUSP
scheme161 is implemented. For the present analysis, a set of 3D non-uniform grids have been
generated assuming the flow symmetry with respect to the passage middle plane. All the grids are of
3-block H-O-H structure (see Figure 6.21). Each mesh covers one half of the blade channel height,
and was obtained by translation of a 2D grid along the span wise direction. Grids of the best
resolution have 45 nodes along this direction, clustered to the end-wall. The distance from the first
cell center to the end-wall was equal to 0.2×10-4 x Сx that produced the area averaged yp + value of 0.8.
As a result of special computations, it has been established that a further grid refinement with respect
to the span wise direction is not necessary.
Below the main attention is paid to the effects of grid refinement in the planes parallel to the end-
wall (in fact, a starting 2D grid refinement), especially in the leading edge region where 3D vortex
structures arise. In order to characterize the grid quality in the LE region, we have introduced an
averaged cell size, Δ*, evaluated for the cells that
are placed in the middle between the saddle Mesh Cell # Δ*/Cx
(separation) point and the blade LE, except the A 360000 0.027
boundary layer region adjacent to the blade. Note
that this cell size is measured in the line of the LE, B 730000 0.022
and the cell aspect ratio in this region did not
exceed 2.0 for all the grids. Table 6.1 covers data C 750000 0.022
for five of the grids, results for which are presented D 760000 0.017
below. For the finest grid, the averaged cell size
introduced is of 1% of the blade axial chord (or E 1200000 0.010
about 5% of the blade LE radius). Note also that the
grid slice given in Figure 6.22 corresponds to grid Table 6.1 Parameters of the Grids Used
B (every second grid line of the grid is omitted
there for clarity).
To ensure a direct comparison of computational heat transfer results with the measurement data,
local Stanton numbers were calculated using the same procedure as developed and justified by [Giel
et. al]162. Remarkably that this procedure results in Stanton numbers that are practically
independent of the surface heat flux rate varied in the experiments. Under the operation conditions
under consideration, the choice of temperature difference used as the driving potential and the
choice of a reference temperature for gas thermos physical properties significantly affects the heat
transfer coefficient. Having performed a comparison of various definitions, [Giel et. al]163 suggested
to define the Stanton number as follows

161 Jameson, A., “Positive schemes and shock modelling for compressible flows”, Int. J. Num. Meth. Fluids, Vol. 20,
743-776 (1995).
162 Giel, P.W., Thurman, D.R., Van Fossen, G.J., Hippensteele, S.A, and Boyle, R.J., “End-wall heat transfer

measurements in a transonic turbine cascade,” ASME Paper 96-GT-180 (1996).


163 Giel, P.W., and Gaugler, R.E., “NASA Blade 1. End-wall heat transfer data. Version 1,” NASA-Glenn Research

Center, Turbine Branch, CD ROM (2001).


146

qw
St =
ρref Uin Cp (Tw − Taw )
Eq. 6.3
where the local adiabatic wall temperature, Taw, is defined as

Taw 1−r
=r+ 2
Tin 1 + 0.5(Υ − 1)Mis
Eq. 6.4
In expression Eq. 6.4 the local isentropic Mach number, Mis, is determined from the wall static
pressure, and the recovery factor, r, is evaluated as r = Pr1/3. The density, ρref, included in Eq. 6.3,
unlike the value used for the Reynolds number is not the actual physical density. It is defined as

pin
ρref = , Tref = Tis + 0.5(Tw − Tis ) + 0.22(Taw − Tis )
RTref
Eq. 6.5
The isentropic temperature, Tis, is evaluated using the local isentropic Mach number and the free-
stream total temperature.
6.10.5.3 Results and Discussion
Flow field computed is illustrated in Figure 6.22. Here, as an example, results obtained with grid B
and the SST version of the Menter model (M-SST) are given. It should be emphasized, however, that
for the Mach number field and the static pressure distribution over the blade all other combinations
of grids A to E and the turbulence models (introduced above) produced practically identical results.
Figure 6.22 shows mid-span Mach
number contours. The distributions
computed are in an excellent agreement
with the NASA GRC experimental data. The
strong affection of the span wise location on
the static pressure distribution over the
suction surface is well reproduced in the
CFD-analysis. Near the trailing edge the
computations predicts an increase in
pressure on the suction side that is due to
the flow overexpansion and viscous-
inviscid interaction phenomena in the
trailing edge region. In contrast to the blade
surface pressure distribution,
computational results for local end-wall
heat transfer are very sensitive both to the
turbulence model and the grid quality.
Figure 6.23 presents Stanton number
distributions over the end-wall computed Figure 6.22 Computed Mid-Span Mach Number
Distribution
with three turbulence models in
comparison with the measurement data.
These distributions were obtained using grid B. Generally one can conclude that all the models
capture the main trends in formation of the local heat transfer pattern under action of 3D vortex
structure developing in the blade passage. However, there are pronounced distinctions in the blade
LE region where spots of enhanced heat transfer are observed. In particular, the k - ω model predicts
147

Figure 6.23 End-Wall Stanton Number (103) Distributions Computed with Grid B in Comparison
with the Measurement Data: (1) k-ω Turbulence Model, (2) M-SST, (3) v2-f, (4) Experiment

a zone of the highest Stanton numbers that adjoins directly the leading edge, but the latter is in
contradiction with the measurements. Both the M-SST model and the v2-f model predict crescent
zones of extreme heat transfer placed slightly upstream of the blade LE. Such a crescent zone is seen
in the experimental Stanton number distribution as well. At that, the v2-f model gives a Stanton
number distribution that is much more non-uniform as compared with the measurement data. All
the models did not predict a spot of enhanced heat transfer observed in the experiments at the blade
suction side not far from the leading edge. As a whole, the MSST model has prediction superiority
among the turbulence models examined.
6.10.5.4 End-Wall Heat Transfer Sensitivity w.r.t Grid and Turbulence Models
Now we concentrate at analysis of grid-sensitivity of local end-wall heat transfer. This analysis is
performed for the k-ω and the M-SST turbulence models taking into account that currently they are
rather popular in predictions of wall-bounded flows. Our computations have shown that the k-ω
model prediction results are considerably less sensitive to grid refinement as compared with the
MSST model. With the k-ω model, grids B to E produced practically identical results. Small
distinctions in the end-wall Stanton number distributions are observed only when comparing results
148

obtained using the initial grid (grid A) with those of one of the finer grids as we discuss the reasons
of relatively week grid-sensitivity of the k-ω model results for the problem under consideration.

Figure 6.24 Effect of Grid Refinement on the End-Wall Stanton Number (x103) Prediction with the
M-SST Turbulence Model: (1) Grid С, (2) Grid D, (3) Grid E, (4) Experiment.

Figure 6.24 illustrates the effect of grid refinement on the end-wall Stanton number distributions
computed with the M-SST turbulence model (in order to sharpen the effect, in this figure the upper
limit of the color legend is decreased as compared with Figure 6.23). The simulation results are very
sensitive to the grid quality in the region placed upstream of the blade LE, where horseshoe vortex
structures arise. Grid refinement results in formation of two crescent zones of extreme heat transfer
(see the St maps for grid D and E), whereas only one such a zone was observed when using coarser
grids. Remarkably that even for grids C and D, comprised of about same numbers of cells, there is a
considerable difference between the results for zones of high Stanton numbers. Grid E produces the
most detailed pattern of the Stanton number distribution. It should be recognized however that the
grid refinement has not resulted in a considerably better agreement between the computational and
measurement data. In particular, as in the coarser grid case, the finest grid computations do not
149

predict the high-St spot observed in the experiments at the blade suction side not far from the leading
edge. As well, the St values are underestimated at the pressure side near the blade trailing edge.
Let’s discuss now the reasons of considerable distinctions between the end-wall heat transfer
prediction results obtained for the blade LE region with the k-ω turbulence model and the M-SST
model. A detailed analysis of the flow structure in the end-wall boundary layer just upstream of the
blade LE has shown that the k-ω model produces a much simpler flow topology than the M-SST
model, provided fine grids are used in both the cases. For the M-SST model case, Figure 6.25 (A)
presents a pattern of limiting streamlines on the end-wall computed with grid E. One can see trails of
a number of well-resolved vortex structure determining the end-wall heat transfer peculiarities in
the blade cascade under consideration. The section A-A marked in the blade LE region was used for
visualization (see Figure 6.25 (B)) of the near-end-wall flow topology in the normal plane. With a
fine grid, the MSST model predicts a complicated vortex structure: with the main horseshoe vortex,
a counter rotating secondary vortex located closer to the end-wall and a tertiary vortex. Application
of a similar visualization technique to the flow field computed with the k-ω model has shown that a
structure with one horseshoe vortex is predicted, and intensity of this vortex is reduced as compared
with the main vortex predicted by the M-SST model.

(A) End wall streakline visualization

(B) Streamline topologies in plane A-A

Figure 6.25 End-Wall Streak Line Visualization

Previously, such a kind of distinctions was reported by [Levchenya et al]164 when analyzing numerical
simulation results for the 3D turbulent flow and end-wall heat transfer in a cascade of thick vanes. In
that contribution it was reported also that, at least in the region of the horseshoe vortex formation,
the k-ω model produces a higher level of the eddy viscosity than the M-SST model, and it is a main
reason of distinctions in the flow topology upstream of the blade leading edge.
6.10.5.5 Summary
With an finite-volume Navier-Stokes code of second-order accuracy, effects of computational grid
refinement have been investigated. The problem of 3D turbulent flow and end-wall heat transfer in

164Levchenya, А.М., Ris, V.V., and Smirnov, E.M., “Testing of turbulence models as applied to calculations of 3D
flow and end-wall heat transfer in cascades of thick vane blades”, Proc. 4th Russian National Heat Transfer Conf.,
MPEI Publishers, Moscow, Russia, Vol.2, 167-170 (in Russian, 2006)
150

a linear transonic turbine cascade with a large turning angle. Three turbulence models were used at
the computations (k-ω model by Wilcox, Menter SST model, v2-f model). The main attention for the
grid-sensitivity aspects was paid to the cases of the k-ω and the M-SST since currently they are rather
popular in predictions of wall bounded flows. It has been established that the M-SST model prediction
results are considerably more sensitive to grid refinement as compared with the k-ω model,
especially for the flow and heat transfer region placed upstream of the blade leading edge, where
horseshoe vortex structures arise. A less grid-sensitivity of the k-ω model is due to the fact that
generally it produces a higher level of the eddy viscosity, and it results in prediction of a simplified
flow topology as compared with the M-SST model165. Other similar studies were performed by
[Ivanov et al.]166 and [V.D. Goriatchev, et al.]167.
6.10.6 Case Study 2 - Comparison of Steady and Unsteady RANS Heat Transfer Simulations of Hub
and End all of a Turbine Blade Passage
The necessity of performing an unsteady simulation for the purpose of predicting the heat transfer
on the end wall surfaces of a turbine passage is addressed by [El-Gabry and Ameri]168. This is
measured by the difference between the two solutions obtained from a steady simulation and the
time average of an unsteady simulation. The heat transfer coefficient (Nusselt number) based on the
adiabatic wall temperature is used as the basis of the comparison. As there is no film cooling in the
proposed case, a computed heat transfer coefficient should be a better measure of such difference
than, say, a wall heat flux. Results show that the effect of unsteadiness due to wake passage on the
pressures and recovery temperatures on both hub and casing is negligible. Heat transfer on the end
walls, however, is affected by the unsteady wake; the time-averaged results yield higher heat
transfer; in some regions, up to 15% higher. The results for the end wall heat transfer were compared
with results in open literature and were found to be comparable.
6.10.6.1 Introduction
Gas turbine heat transfer remains an important topic of concern as turbine inlet temperatures
continue to rise and combustor exit profiles continue to flatten with the goal of maximizing power
output and efficiency. This in turn means higher gas-side heat transfer to the hot gas path
components, including the vane and blade and, in particular, the end walls, which are now
experiencing temperatures nearly as high as the peak temperatures near the midspan.
Of importance to understanding the end wall heat transfer is describing and characterizing end wall
flows. As early as 1976, [Langston et al.]169 wrote that “the literature is certainly not lacking in
experimental studies of end wall flows in turbine cascades” but is lacking a “complete analytical
solution” of the end wall flows. In this 1976 paper, [Langston et al. ] cited the potential of numerical
models to fill in that gap, which, one can argue, has yet to be entirely filled, which, that is, where the
present research and other papers on computational fluid dynamics (CFD) modeling of end wall flow

165 Alexander M. Levchenya* and Evgueni M. Smirnov, “CFD-Analysis of 3d Flow Structure and End-wall Heat
Transfer in a Transonic Turbine Blade Cascade: Effects of Grid Refinement”, West-East High Speed Flow Field
Conference 19-22, November 2007.
166 Nikolay Ivanov, Vladimir Ris, Evgueni Smirnov, and Denis Telnov, “Numerical Simulation Of End-wall Heat

Transfer In A Transonic Turbine Cascade”, Conference On Modelling Fluid Flow (CMFF’03) The 12th
International Conference On Fluid Flow Technologies Budapest, Hungary, September 3-6, 2003.
167 V.D. Goriatchev1, N.G. Ivanov2, E.M. Smirnov2, V.V. Ris2, “CFD Analysis of Secondary Flows and Pressure Losses

in a NASA Transonic Turbine Cascade”, 1- Department of Mathematics, Tver State Technical University, 170026
Russia; and 2- Department of Aerodynamics, St.-Petersburg State Polytechnic University, 195251, Russia.
168 Lamyaa A. El-Gabry, Ali A. Ameri, “Comparison of Steady and Unsteady RANS Heat Transfer Simulations of

Hub and End wall of a Turbine Blade Passage”, Journal of Turbomachinery, JULY 2011.
169 Langston, L. S., Nice, L. M., and Hooper, R. M., 1976, “Three Dimensional Flow Within a Turbine Cascade

Passage,” ASME Paper No. 76-GT-50.


151

and heat transfer continue to play an important role in our understanding of this critical region of
the turbine passage.
[Langston et al. ] used ink to visualize flow near the end walls of a cascade and measured pressures
and velocities at axial locations within the passage. [Gregory-Smith et al.]170 presented the flow
visualization of the end wall and made measurements in a cascade with the goal of calculating the
vorticity in the passage.
There have also been several review papers that survey literature on secondary flows in turbine
passages and that highlight results of turbine end wall aerodynamics and heat transfer studies. The
dominant secondary flow structures highlighted include the cross-passage flow that crosses from the
pressure side to the suction side and what is commonly referred to as the horseshoe
vortex that forms at the stagnation point on the leading edge, where the flow separates into a
pressure side and a suction side leg. There are additional forms of secondary flow, including corner
vortices. There are three “corners”: the first is between the pressure side surface and the hub, the
second is between the suction side surface and the hub, and the third is between the leading edge
surface and the hub.
Near the blade tip, the dominant secondary flow feature is the tip leakage vortex that manifests itself
along the suction side of the airfoil. The losses due to tip leakage flow can account for up to 1/3 of the
total stage losses, as suggested by [Boyle et al.]171, making them important from an aerodynamics
perspective as well as their impact on heat transfer. Research in the topic of near-tip flows and heat
transfer started with basic research using very basic geometry to represent flat and grooved
rectangular tip models. These studies, among other things, established that the effect of relative
motion between the blade and casing on heat transfer is negligible, which was consequently followed
by a series of linear cascade tests using three-blade, four-blade, and five-blade cascades to measure
static pressure and local heat transfer distribution on the airfoil, tip, and casing surfaces for a variety
of tip geometries.
Experimental data obtained in stationary cascades offer detailed measurements; however, results
may not scale to the actual engine conditions or be truly representative of a true rotating blade in an
engine. Therefore, experimental data have also been obtained at engine conditions in rotating rigs.
[Haldeman and Dunn]172 at the Ohio State University Gas Turbine Laboratory measured the heat
transfer for the vane and blade of a rotating high pressure turbine stage operating at design corrected
conditions using a large shock-tunnel facility.
For the blade, Stanton number is reported at 20% and 96% spans at the blade tip and on the shroud.
[Polanka et al.]173 also made pressure and heat flux measurements on the tip and shroud under
rotating conditions at the U.S. Air Force Turbine Research Facility, which is a full scale rotating rig.
These test setups are far more representative of engine conditions, however, the test data are sparse
and a limited number of discrete measurements are available as compared with the cascade tests.
In addition to the experimental research on end wall aerodynamics and heat transfer, there have
been a number of computational studies on the subject, several of which complement the
experimental work. The 9H tip heat transfer experiments of [Bunker et al.]174 were modeled using in-

170 Gregory-Smith, D. G., Graves, C. P., and Walsh, J. A., 1988, “Growth of Secondary Losses and Vorticity in an Axial

Turbine Cascade,” ASME J. Turbomachine, 110, pp. 1–8.


171 Boyle, R. J., Haas, J. E., and Katsanis, T., 1985, “Predicted Turbine Stage Performance Using Quasi-Three-

Dimensional and Boundary Layer Analyses,” J. Propulsion Power, 1, pp. 242–251.


172 Haldeman, C. W., and Dunn, M. G., 2004, “Heat-Transfer Measurements and Predictions for the Vane and Blade

of a Rotating High-Pressure Turbine Stage,” ASME J. Turb., 126, pp. 101–109.


173 Polanka, M. D., Hoying, D. A., Meininger, M., and MacArthur, C. D., 2003, “Turbine Tip and Shroud Heat Transfer

and Loading—Part A: Parameter Effects Including Reynolds Number, Pressure Ratio, and Gas-to-Metal
Temperature Ratio,” ASME J. Turb., 125, pp. 97.
174 Bunker, R. S., Bailey, J. C., and Ameri, A., 2000, “Heat Transfer and Flow on the First-Stage Blade Tip of a Power

Generation Gas Turbine: Part 1—Experimental Results,” ASME J. Turb., 122, pp. 263–271.
152

house Reynolds-averaged Navier–Stokes _RANS_ codes and the OSU tests were modeled using the
commercial CFD code STARCD _17_ and an in-house GE-developed code Tacoma . [Polanka et al.]175
presented a comparison between numerical predictions and experimental results of [Polanka et al.].
The conclusions were that the agreement between the test and the CFD is “very good” on the airfoil
and “moderately good” on the end walls and tip; the tip heat transfer was overpredicted by the CFD
and although the shroud pressure distribution was accurately predicted with the 3D RANS solver,
the heat transfer there was overpredicted by a factor of 2.
Wake effects and unsteadiness were investigated by [Pullman]176, who used test measurements to
show the vortex formation at the stator exit and their development within the rotor passage to the
rotor exit. Using the experimental measurements at the stator exit as boundary conditions to a rotor
CFD model, the predictions of the aerodynamic losses _using steady and unsteady analyses are
calculated and the flow field at the rotor exit is predicted and compared with test data. The steady
simulation predicted 10% less loss in the rotor row than the unsteady simulation.
The present work focuses on the effect of unsteadiness due to wake passage on the end wall heat
transfer as well as pressures and
temperatures as predicted using an
unsteady 3D RANS solver. The vehicle
for this computational study is the E3
high pressure turbine blade, which has
been reported extensively in literature.
Time-resolved test measurements are
not available for the E3 HPT blade;
therefore, qualitative comparisons of the
CFD results will be made with HPT data
in open literature, to serve as a check on
the CFD results. The primary question
this research seeks to answer is whether
an unsteady analysis gives the same
answer as a steady analysis. For this size
model ∿2x106 nodes, it takes 500 clock
hours to reach a converged steady state
solution. Also, it takes the same model
25,000 clock hours to converge the
unsteady simulation. The question is: Is
the added time and complexity of Figure 6.26 Grid on the Solid Surfaces of the Geometry
running unsteady worth it?
6.10.6.2 Computational Method
A 3D RANS code developed at the NASA Glenn Research Center has been used to predict pressures,
temperatures, and heat transfer in a single blade passage of the GE-E3 gas turbine. The code was
Glenn-HT and was described in detail by [Ameri et al.]177. It uses a finite volume discretization
scheme and is second order accurate in time and space. The turbulence model used in the calculations
was the low Reynolds number k-ω model by [Wilcox], which is integrated to the wall; a

175 Polanka, M. D., Clark, J. P., White, A. L., Meininger, M., and Praisner, T. J., 2003, “Turbine Tip and Shroud Heat
Transfer and Loading Part B: Comparison Between Prediction and Experiment Including Unsteady Effects,” ASME
Paper No. GT2003-38916.
176 Pullman, G., 2004, “Secondary Flows and Loss Caused by Blade Row Interaction in a Turbine Stage,” ASME

Paper No. GT2004-53743.


177 Ameri, A. A., Steinthorsson, E., and Rigby, D., 1998, “Effect of Squealer Tips on Rotor Heat Transfer and

Efficiency,” ASME J. Turbomachine., pp. 753–759.


153

nondimensional grid spacing y+ of near 1 is maintained on all wall surfaces, including the blade and
end walls. The code was run on 48 processors of a Xeon Linux Cluster using message passing interface
for parallel processing.
The E3 gas turbine has 46 vanes and 76 blades and the blade rotates at 8400 rpm. The blade tip
clearance is 2% of the blade span. The present research focuses on the effect of the upstream vane
wake on the blade. The domain of this computation is restricted to a single blade passage; this
simplification of using a vane/blade ratio of 1 is based on a separate preliminary study, which showed
that for purposes of computing average heat transfer, the wake frequency for a 1:1 ratio produces a
very similar result as compared with a 2:3 ratio. The 2:3 ratio is an approximation to the actual
vane/blade count, which was 46:76.
The inlet of the domain is located at 15% axial chord upstream of the blade leading edge _midway
between the vane trailing edge and the blade leading edge_ and the exit of the domain is at 50% axial
chord downstream of the blade trailing edge. To create the multiblock structured grid needed for the
solver GRIDPRO™, commercially available software for generating structured meshes was used.
shows the surface mesh on the solid airfoil and hub. Figure 6.27a shows the grid on the casing
surface highlighting the multiblock structure of the grid blocks. Figure 6.27b shows the grid
structure on the hub surface. The grid consists of 164 blocks and a total of 1.8 x 106 nodes. There are
65 nodes across the tip clearance gap in the radial direction and 101 nodes from the hub to tip of the
blade. The grid independence of the solution was established.

Figure 6.27 (a) Casing surface mesh showing multiblock structure and (b) hub surface mesh showing
multiblock structure

A dimensionless time step of 0.005 was used based on earlier investigations in which the time step
was varied from 0.001 to 0.01. For the selected time step of 0.005, 320 steps were required to
complete one period; (i.e., the passing of a wake across a single blade passage), which is sufficiently
fine, a resolution for purposes of this study. Based on the selected time step, blade count, and RPM,
there were 320 time steps over the period of one wake passage.
6.10.6.2.1 Boundary Conditions
A separate computation of the vane flow was used to establish the inlet boundary conditions used in
the present blade analysis. The total pressure and temperature at the vane exit were calculated using
the same code and methodology and the wake profile is taken from the midspan of the vane
154

computation and used for the entire span of the inlet section to the blade. The total pressure in the
vane wake was approximated with a trigonometric function

pt,θ (0, θ) = p0,bg {1 − 0.15sin[nθ/2 + πt/τ]10 }


Eq. 6.6
Likewise, the total temperature wake behind the vane was approximated using a second
trigonometric function

Tt,θ (0, θ) = T0,bg {1 − 0.05sin[nθ/2 + πt/τ]10 }


Eq. 6.7
Figure 3 shows the normalized inlet total temperature and total pressure wakes as applied to the
blade inlet. The wake turbulence and length scale were similarly obtained and approximated with
trigonometric functions and specified at the blade inlet. The background level for turbulence
intensity was 2% and the amplitude was 5% for a peak value of 7%, i.e.,

Tu(t, θ) = Tub,g + Tuamp {sin[nθ/2 + πt/τ]6 }


Eq. 6.8
The wake turbulence length scale was also fitted with the same function, where the background
length scale was set to 10% axial chord and the peak length scale was set to 25% axial chord. To
simulate the near wall boundary layer effects, a boundary layer profile for the inlet total pressure
and temperatures is applied at the inlet section that extends to 1% span from the hub and casing. At
the exit of the domain, a constant static pressure boundary condition is applied at the hub and a radial
equilibrium is enforced. No-slip boundary conditions are applied to all solid surfaces in the domain,
including airfoil, hub, and casing. The heat transfer coefficient is defined as

Qw
h=
Taw − Tw
Eq. 6.9
where Qw is the wall heat flux, Tw is the wall
temperature, and Taw is the adiabatic wall
temperature. Therefore, in order to calculate heat
transfer coefficient defined as such, it is necessary to
run two separate analyses; one in which the walls
have zero heat flux in order to determine the
adiabatic wall temperatures and a second with a
prescribed wall temperature in order to calculate the
heat flux. For the runs with a prescribed wall
temperature, a constant temperature of 0.7
(normalized using inlet total temperature) is applied Figure 6.28 Total temperature (T0) and
to all walls. The steady state boundary conditions are total pressure (P0) at the blade inlet
based on the average of the unsteady computations.
6.10.6.3 Results and Discussion
Figure 6.30 shows the instantaneous pressure distribution on the hub surface at equal times over
the period of one wake passage. The image outlined with the dashed line shows the time-averaged
pressure distribution. These time-averaged results are compared with a steady simulation in which
there is no wake at the inlet. The inlet boundary condition for the steady simulation is the area
average of the unsteady boundary condition in Figure 6.28.
155

Figure 6.29 Instantaneous and time-averaged (dashed) hub heat transfer distribution for a wake
passing

The instantaneous pressure distributions in Figure 6.30, particularly the position and extent of the
low pressure region on the suction side, illustrate the effects of the wake passage. This region of
minimum pressure is indicative of a vortex that is traveling along the suction side (perhaps the
suction side (SS) leg of the horseshoe vortex). As time passes from t=0 to t=T/4, this low pressure
region moves along the SS edge and extends in the blade to blade direction. As it approaches the
trailing edge (TE), it appears to move away from the SS edge toward the pressure side (PS) trailing
edge and begins to diffuse. At some time between T/4 and 3/8T, a second low pressure region
emerges near the leading edge before the high point of curvature (high-C) that also moves along the
SS edge extends and diffuses as before. The periodic growth, diffusion, and movement of this low
pressure region indicate an interaction between the end wall secondary flows and the upstream
wake.
Likewise, the instantaneous heat flux images in Figure 6.29 suggest an interaction between the
wake passage and the end wall heat transfer on the hub. In discussing the pressure, we focused on
the low pressure region along the suction side and noted the presence of two low pressure regions
at the same time during part of the period: one region is diffusing into the passage near the trailing
156

Figure 6.30 Instantaneous and time-averaged „dashed… hub surface


pressure for a wake passing

edge and one is beginning to grow as it moves along the SS edge near the leading edge. In Figure
6.29, let us focus our attention on the high heat transfer regions shown in red. At the leading edge,
there is high heat transfer due to stagnation that extends along the PS edge and grows. There is a low
heat transfer region between the passages that moves along the SS edge and grows then diffuses
across the passage to PS edge. In doing so, it mixes with the high heat transfer fluid and reduces the
heat transfer, resulting in what appears to be a break in the high heat flux zone along the PS edge.
The latter piece of the high heat transfer region along the PS edge shrinks in extent and decreases in
magnitude as it diffuses and appears to “shed” at the trailing edge.
The pressure distribution on the casing and hub surfaces is plotted in Figure 6.31 and Figure 6.32,
respectively, nondimensionalized by the inlet total pressure. Figure 6.32 is actually a repeat of the
time-averaged contour plot in Figure 6.29 outlined with a dashed line. The airfoil outline shown in
Figure 6.31 and Figure 6.34 and other casing images to follow is the projection of the tip surface
on the casing.
On the casing surface, there is a region of high pressure in the forward part of the passage above the
pressure side surface of the airfoil whose outline is shown in Figure 6.31. This high pressure
corresponds to the stagnation region of the airfoil near the tip. There is a low pressure region near
the trailing edge of the airfoil, which likely corresponds to higher tip flow speeds.
157

On the hub surface


Figure 6.32, the
stagnation region at the
leading of the airfoil
results in a high
pressure zone that
extends into the
passage along the
pressure side. On the
suction side is a low
pressure region just
past the high-C
_curvature_ of the
airfoil that increases
along the passage likely
to correspond with the
suction side leg of
horseshoe vortex
forming at the leading
edge and growing as it Figure 6.31 Time-Averaged Casing Pressure
moves through the
passage and interacts
with the cross-passage flow.
Contour plots of pressure on the casing and hub surfaces were also obtained from a steady state CFD
model and compared with the time-averaged distributions. There was no discernable difference
between the time-
averaged and steady
state contours that
could be identified from
the contour plots.
Therefore, the
difference between the
time-averaged and
steady state local
pressures is plotted
instead in Figure 6.34
and Figure 6.33.

Figure 6.32 Time-Averaged Hub Pressure


158

Figure 6.34 shows


the difference
between the time-
averaged and steady
state pressure results
on the casing surface
as a percentage of the
time-averaged results.
Clearly, there is
negligible difference
between the two
solutions; the largest
difference is on the
order of 2% and is
located slightly
forward of the high-C
point on the casing
surface and further aft
at the high-C point on
the hub surface. Figure 6.34 Difference in casing pressure distribution between the time-
However, a 2% averaged and steady results
difference is not
significant enough to
justify/warrant a transient CFD analysis.
Figure 6.35 and Figure 6.36 show the
adiabatic wall temperature distribution
on the casing and hub surfaces,
respectively, nondimensionalized by the
inlet total temperature. Note that the
two figures are not plotted at the same
scale to illustrate trends and enable
discussion. In Figure 6.35, the adiabatic
wall temperature on the casing surface
exceeds the inlet total temperature (i.e.,
the dimensionless Taw exceeds 1); this
occurs at the stagnation region and along
the pressure side edge of the airfoil tip.
The rise in adiabatic wall temperature
Figure 6.33 Difference in hub pressure distribution
can be attributed to work processes in between the time-averaged and steady results
the near-tip region, where energy is
imparted from the rotating fluid in the
tip gap to the stationary casing, thereby increasing the total temperature at the casing surface.
The rise in adiabatic wall temperature on the casing surface was noted by [Thorpe et al.]
experimentally and numerically: “In the tip gap, the stagnation temperature is shown to rise above
that found at stage inlet by as much as 35% of stage total temperature drop.” A similar finding is
ascertained from Figure 6.35, which shows the maximum adiabatic wall temperature to be 1.06 and
the exit adiabatic wall temperature to be 0.84; therefore, the rise in adiabatic wall temperature 0.06
is about 1/3 of the total temperature drop for the rotor passage.
On the hub surface Figure 6.36, there are no regions where the dimensionless adiabatic wall
temperature exceeds unity since there are no rotating/stationary frames of references in the hub
159

wall unlike the casing. The


adiabatic wall temperature
decreases along the flow
passage because work is
extracted from the stage.
The low pressure region of
Figure 6.32 discussed
earlier as being
attributable to
recirculation in the horse
shoe vortex results in a
region of minimum
adiabatic wall
temperatures at the high-C
point and aft. The
adiabatic wall temperature
results in Figure 6.35 and
Figure 6.36 are obtained
through the time-
averaging of the transient Figure 6.35 Time-averaged casing adiabatic wall temperature
CFD solutions and are
compared with steady
state results. Figure 6.37 and
Figure 6.38 use the same scale and
show the difference between the
time-averaged and steady state
adiabatic wall temperatures as a
percentage of the temperature
difference between the adiabatic
wall and the isothermal wall, i.e.,

̅aw − Taw
T
̅aw − Tw
T
Eq. 6.10
This is a more physical
representation of the difference as Figure 6.36 Time-averaged hub adiabatic wall temperature
the denominator is indicative of the
driving temperature for heat
transfer. Figure 6.37 shows that the time-averaged casing surface adiabatic wall temperatures are
about 4-6% lower than the steady results. For the hub surface, Figure 6.37 shows that the time-
averaged hub surface adiabatic wall temperatures are up to 8% lower than those predicted by a
steady state simulation. The region of largest difference is in the forward region of the hub surface.
160

Figure 6.37 Difference in casing adiabatic wall temperature distribution between the time-averaged and
steady results

The dimensionless wall heat flux, as defined in the nomenclature, is plotted in Figure 6.39 and
Figure 6.40. Figure 6.39 shows the time averaged wall heat flux distribution on the casing surface.
The area of highest heat transfer is the region of the casing above the pressure side edge of the airfoil
tip due to the entry of tip leakage flow into the clearance gap, consistent with Figure 6.35.
The heat transfer rates at the hub surface shown in Figure 6.40 repeat of the dash-outlined image
in Figure 5.30 are lower than those found on
the casing surface Figure 6.39. The area of
highest heat transfer is on the hub surface
near the leading edge in the stagnation region
and penetrates at a lower magnitude along
the pressure side surface aft toward the
trailing edge. The heat transfer rate on the
end walls obtained through time averaging is
compared with the steady state-predicted
heat fluxes and the difference between the
time-averaged and steady results for the
casing and hub are shown in Figure 6.41 and
Figure 6.42, respectively.
Figure 6.38 Difference in hub adiabatic wall
temperature distribution between the time-averaged
and steady results
161

On the casing surface Figure 6.41,


the time-averaged solution predicts
up to 10% higher heat transfer in
some regions, specifically in the
region above the SS edge of the tip.
The streaks of large positive and
large negative differences past the
trailing edge in the wake of the blade
can be neglected; they likely indicate
a slight offset in the distribution
rather than a fundamental difference
in the prediction.
On the hub surface Figure 6.42, the
unsteadiness increases the heat
transfer by up to 10%; the largest Figure 6.39 Time-averaged casing heat transfer rate
difference is at the leading edge and
along the pressure side edge into the passage. Generally, there are larger differences in the local heat
transfer prediction on the hub surface than on the casing surface with the effect on the casing being
highly localized to the region past the suction side and nearly zero difference everywhere else.
Figures 18 and 19 of [El-Gabry and Ameri]178, shows the Nusselt number distribution on the casing
and hub surfaces, respectively, while Figure 6.43 show the difference between the time-averaged
and steady state Nusselt numbers.
The region of highest Nusselt
number on the casing surface is also
the region of highest heat transfer
Figure 6.39 and highest adiabatic
wall temperature Figure 6.35,
which is above the pressure side
edge of the blade tip. Not
surprisingly, the region showing
significant difference in the effect of
time-averaging is at the same
location, where there were
significant differences in the heat
flux predictions Figure 6.41; the
effect of the wake on adiabatic wall
temperature for the casing surface Figure 6.40 Time-averaged hub heat transfer rate
was shown to be negligible Figure
6.37.

Lamyaa A. El-Gabry, Ali A. Ameri, “Comparison of Steady and Unsteady RANS Heat Transfer Simulations of
178

Hub and End wall of a Turbine Blade Passage”, Journal of Turbomachinery, JULY 2011.
162

On the hub surface, the region of


highest Nusselt number is at the
stagnation region along the airfoil
leading edge, as shown in figure 19
of [El-Gabry and Ameri]179, this was
also the region of highest heat flux
per Figure 6.40. Whereas the
casing surface adiabatic wall
temperatures were unaffected by
the wake; the hub surface
temperatures were lower when the
unsteadiness due to wake passage
was considered. The heat flux
distribution on the hub was also
influenced by the wake passage.
Therefore, one finds that the Nusselt
number, which is driven by heat flux
Figure 6.41 Difference in casing heat transfer rate between
and temperature difference, is the time-averaged and steady results
influenced by wake passage, as
shown.
The largest effect unsteadiness due
to wake passage has on Nusselt
number is along the leading edge of
the airfoil but also along the
pressure side surface. The
unsteadiness increases the local
Nusselt numbers by up to and above
10%.
6.10.6.4 Heat Transfer
Comparison With Open
Literature
As previously mentioned, there are
no experimental results for end wall
pressures, temperatures, or heat
flux for the GE-E3 HPT blade.
However, there are other results for
other engine blades in open
literature that can serve as a
qualitative comparison to ensure Figure 6.42 Difference in hub heat transfer rate between the
that the trends predicted by the CFD time averaged and steady results
are sensible.

Lamyaa A. El-Gabry, Ali A. Ameri, “Comparison of Steady and Unsteady RANS Heat Transfer Simulations of
179

Hub and End wall of a Turbine Blade Passage”, Journal of Turbomachinery, JULY 2011.
163

6.10.6.4.1 Comparison of Casing Heat Flux With [Epstein et al.]


[Epstein et al.]180 measured the time-resolved static pressure and heat flux to the casing wall in a
blow down turbine test rig. The turbine tested was a transonic high pressure turbine scaled from a
Rolls Royce engine. A detailed quantitative comparison with the casing measurements of Epstein et
al. can be found in Ref.181. A key finding of that research is that a large percentage of the total casing
heat flux, nearly 45% of the total heat
load, comes from the flow over the
rotor blade tip; the tip area
represents about 30% of the total
casing area and contributes 45% of
the total heat flux. Similarly, the
present results for the E3 show that
the above rotor tip region constitutes
about 12% of the total casing surface
area and contributes about 22% of
the total heat casing heat load.
A comparison of Fig. 22 (Not shown)
viewing the experimental results for
the casing heat flux on data from Ref.
182and Figure 6.39 showing the

CFD-predicted nondimensional heat


flux for the E3 show similar trends;
the region of high heat transfer is in
the over tip region above the
pressure side edge and decreases Figure 6.43 Difference in hub Nusselt number between the
toward the trailing edge. A region of time averaged and steady results
slightly higher heat flux extends from
the leading edge toward the inlet. The heat flux decreases along the passage and reaches a minimum
at the rotor wake region.
6.10.6.4.2 Comparison of Hub Heat Transfer With Tallman et al.
[Tallman et al.]183 presented contour plots of the hub Stanton number using a RANS solver. Figure 23
of [El-Gabry and Ameri]184, shows the CFD-predicted Stanton number. The test data are shown in Ref
185; however, the measurements are sparse on the hub surface and insufficient to generate a contour

plot. Therefore, the CFD results will be used for comparison.


Notable features in the Stanton number distribution include the high heat transfer region at the
leading edge and the moderate to high heat transfer along the pressure side edge of the passage.

180 Epstein, A. H., Guenette, G. R., Norton, R. J. G., and Yuzhang, C., 1985, “Time Resolved Measurements of a
Turbine Rotor Stationary Tip Casing Pressure and Heat Transfer Field,” AIAA Paper No. 85-1220.
181 Ameri, A. A., Rigby, D. L., Steinthorsson, E., Heidmann, J., and Fabian, J. C., 2008, “Unsteady Analysis of Blade

and Tip Heat Transfer as Influenced by Upstream Momentum and Thermal Wakes,” ASME Paper No. GT2008-
51242.
182 See 165.
183 Tallman, J. A., Haldeman, C. W., Dunn, M. G., Tolpadi, A. K., and Bergholz, R. F., 2006, “Heat Transfer

Measurements and Predictions for a Modern High Pressure, Transonic Turbine, Including End walls,” ASME Paper
No. GT2006-
90927.
184 Lamyaa A. El-Gabry, Ali A. Ameri, “Comparison of Steady and Unsteady RANS Heat Transfer Simulations of

Hub and End wall of a Turbine Blade Passage”, Journal of Turbomachinery, JULY 2011.
185 R. F., 2006, “Heat Transfer Measurements and Predictions for a Modern High Pressure, Transonic Turbine,

Including End walls,” ASME Paper No. GT2006-90927.


164

[Figure 24 (Not shown here)]186 presents the time-averaged Nusselt number distribution for the hub
surface of the E3 blade; the data is identical to [Fig. 19 (same source)] presented earlier; however,
the range has been modified to better highlight features and draw comparisons to the Stanton
number distribution for the hub in [Fig. 23]. The contours in [Figs. 23 and 24]187 are similar: The
heat transfer peaks at the leading edge due to stagnation and decreases with a relatively high heat
transfer region along the PS edge. There is a much narrower low heat transfer region along the
suction side, similar to the blade analyzed.
6.10.6.5 Conclusions
The end wall pressures, temperatures, and heat transfer have been computed using an unsteady
RANS calculation, time averaged, and compared with steady state predictions to evaluate the effect
of unsteadiness due to wake passage.
Results obtained using the k-omega turbulence model show that the effect of unsteadiness on the
average end wall pressures on both hub and casing is negligible; the unsteadiness increases the
pressure prediction in some areas by up to 2%. The area, where the pressure is most affected by
unsteadiness, is on the suction side part of the passage for both hub and casing and the effect of
unsteadiness seems to extend over a slightly larger area on the hub than the casing.
The contribution of unsteadiness due to wake passage on the recovery temperature on the casing is
negligible over most of the surface. On the hub surface, the time-averaged results yield lower
recovery temperature predictions in the forward region of the passage.
Heat transfer on the end walls is affected by the unsteady wake more than the pressures and
temperatures. The time-averaged results yield higher heat transfer; in some regions, the increase is
up to 15%. On the casing surface, the increase in heat transfer is more localized and appears on the
suction side region of the passage, where the peak pressure difference between the time averaged
and steady results occurred. The effect on the hub surface is less localized and the increase in heat
transfer predictions can be seen all along the pressure side edge and penetrate well into the passage.
As there are no experimental data available to validate the CFD results for the E3, the results for the
end wall heat transfer were compared with results in open literature. For the casing, the results were
comparable to experimental measurements of [Epstein et al.] quantitatively in terms of the influence
of the above-rotor tip area to the overall shroud heat flux and qualitatively in terms of the pattern of
heat transfer on the surface. For the hub surface, the results were compared with results from
[Tallman et al.] and were also found to be similar, thereby verifying the unsteady Reynolds-averaged
Navier-Stokes (URANS) calculations of the present work.

186 See Previous.


187 See Previous.
165
166

7 Blade Cooling
As the turbine inlet temperature increases, the heat transferred to the turbine blade also increases.
The level and variation in the temperature within the blade material, which cause thermal stresses,
must be limited to achieve reasonable durability goals188. The operating temperatures are far above
the permissible metal temperatures. Therefore, there is a critical need to cool the blades for safe
operation. The blades are cooled with extracted air from the compressor of the engine. Since this
extraction incurs a penalty on the thermal efficiency and power output of the engine, it is important
to understand and optimize the
cooling technology for a given turbine
blade geometry under engine
operating conditions. Gas turbine
cooling technology is complex and
varies between engine manufacturers.
Figure 7.1 shows the common cooling
technology with three major internal
cooling zones in a turbine blade as film
cooling in the leading edge, pressure
and suction surfaces with Vanes, and
blade tip region. The leading edge is
cooled by jet impingement with film
cooling, the middle portion is cooled
by serpentine rib-roughened passages
with local film cooling, and the trailing
edge is cooled by pin fins with trailing
edge injection. Interested readers are
referred to several recent publications
that address state-of-the-art reviews
of turbine blade cooling and heat
transfer.
The first high pressure stage of a
modern gas turbine operates at very Figure 7.1 The Schematic of a Modern Gas Turbine Blade
high temperatures that require with Common Cooling Techniques (Courtesy of Je-Chin Han)
complex blade-cooling systems to
guarantee high performance and efficiency of the gas turbine while maintaining a very low level of
energy losses, though using compressed air for cooling189. An accurate and efficient Conjugate Heat
Transfer (CHT) solver is thus necessary to compute the flow and temperature fields of the air within
the cooling channels and of the gas around the blades by means of the Navier Stokes and energy
equations as well as the blade temperature field, by means of the heat conduction equation. Due to
the very high geometrical complexity of the cooling channels within the blades, generating a body
fitted mesh for the three domains air, gas and blade is extremely difficult and time consuming. And
Nevertheless, many turbine blade cooling simulations have been performed with success, though at
large computational cost190.

188 Je-Chin Han, “Recent Studies in Turbine Blade Cooling”, International Journal of Rotating Machinery, 2004.
189 D. De Marinisa, M. D. de Tullioa, M. Napolitanoa,* and G. Pascazioa, “A conjugate-heat-transfer immersed-
boundary method for turbine cooling”, Energy Procedia, 2015.
190 Luo G. and Razinsky E. H., “Conjugate heat transfer analysis of a cooled turbine vane using the V2F turbulence

mode,l” Journal of Turbomachinery, (2007).


167

[Mahmood and Acharya]191 considered adding fillets at the junction of end wall and blade/vane
leading edge in the cascades. The blade cascade operates at low speed atmospheric conditions while
the vane cascade operates at high speed conditions with the exit Mach number near to 1.0.

7.1 Film Cooling Effects


According to the theory of Carnot cycle, increasing the inlet temperature of gas turbine is an effective
way to increase the efficiency and capacity of a turbine. In order to enhance the performance of jet
engine and gas turbine, the temperature of the gas flowing into a turbine blade passage has been
raised continually in recent years, which might result in the damage of blades, especially the leading
edge (LE) which is exposed to the hot gas directly. Although new high temperature materials have
been investigated and used constantly, it is obvious that they couldn’t follow the rising pace of the
inlet temperature.
Four types of cooling methods, such as
➢ Convection cooling,
➢ Impingement cooling,
➢ Film cooling,
➢ Effusion cooling,
and their hybrid methods are used in practical
engineering. Usually, the effusion cooling can
provides the best cooling among these four
methods, while it is seldom used because it will
weaken the structure strength of blades
greatly. Convection cooling and impingement
cooling are usually used in conditions where
the temperature is lower than 1600˚K since
they cannot provide protection to the surface of
blades. Film cooling is the only way can be
used in whole range of the temperature is Figure 7.2 Typical high-pressure turbine stage
higher than 1600˚K. Figure 7.2 [Owen, showing rim seal and wheel-space
2009], illustrates a typical high-pressure gas
turbine stage showing the rim seal and the wheel-space between the stator and the rotating turbine
disc. It is curved downstream under the press and friction of hot main flow colored in pink in the
figure, then forms a thin cooling film on the surface which separates the blade surface from hot gas.
Meanwhile, it takes the sporadic flames and radiant heat to downstream. Hence, it can protect the
blade surface effectively. In the next section we divert our attention to the Secondary Flow which is
another cause of unsteadiness and complication in Turbomachinery.
7.1.1 Fundamentals of Film Cooling
Film cooling primarily depends on the coolant-to hot mainstream pressure ratio (Pc/Pt), temperature
ratio (Tc/Tg), and the film cooling hole location, configuration and distribution on a film cooled airfoil.
The coolant-to-mainstream pressure ratio can be related to the coolant-to-mainstream mass flux
ratio (blowing ratio) while the coolant-to main stream temperature ratio can be related to the
coolant-to-mainstream density ratio. In a typical gas turbine airfoil, the Pc/Pt ratios vary from 1.02 to
1.10, corresponding blowing ratios approximately from 0.5 to 2.0, while the Tc/Tg values vary from
0.5 to 0.85, corresponding density ratios approximately from 2.0 to 1.5. Both pressure ratio (Pc/Pt),
and temperature ratio (Tc/Tg), are probably the most useful measure in quantifying film cooling

191Gazi I. Mahmood, Sumanta Acharya, “Blade and Vane Leading Edge Fillet on End wall Cooling in Linear
Turbine Cascades”, Proceedings of the 15th International Heat Transfer Conference, IHTC-15 August 10-15,
2014, Kyoto, Japan.
168

effectiveness since these ratios essentially gives the ratio of the coolant to hot mainstream thermal
capacitance.
In general, higher the pressure ratio, better the film cooling protection (i.e., reduced heat transfer to
the airfoil) at a given temperature ratio, while lower the temperature ratio, better the film cooling
protection at a given pressure ratio. However, too high of pressure ratio (i.e., blowing too much) may
reduce the film cooling protection, because of jet penetration into the mainstream (jet lift-off from
the surface). Therefore, it is important to optimize the amount of coolant for airfoil film cooling under
engine operating conditions (Reynolds number 106, Mach number0.9 at exit conditions). For
designing a better film cooling pattern of an airfoil, turbine cooling system designers also need to
know where heat is transferred from hot mainstream to the airfoil.
As mentioned earlier, these film-hole pattern (i.e., film hole location, distribution, angle and shape)
would affect film cooling performance. There are many film cooling papers available in the open
literature. This paper is limited to review a few selected papers to reflect recent development in
turbine blade film cooling. Specifically, this review focuses on the following topics: rotor blade film
cooling, nozzle guide vane film cooling, airfoil end wall film cooling, as well as airfoil leading edge and
blade tip region film cooling.
Several papers for rotator blade, stator
vane, and end wall film cooling, under
representative engine flow conditions,
are reviewed and discussed. Some
parametric studies for large-scale
cascade blade, cascade vane, cascade
end wall, simulated blade tips and
leading edge region film cooling under
low speed flow conditions are also
reviewed and discussed.
Figure 7.3 shows the schematic of
film cooling concept. Typically, the
heat load to the surface without film Figure 7.3 Schematic of Flm Cooling Concept
cooling is represented as heat flux q’=
h0(Tg- Tw), where h0 is the heat transfer coefficient on the surface with wall temperature Tw and
oncoming gas temperature (Tg). When coolant is injected on the surface, the driving temperature is
Tf, film temperature, which is a mixture of gas (Tg) and coolant temperature (T), q"= h(Tf- Tw), where
h is the heat transfer coefficient on the surface with film injection. Also, a new term film effectiveness
0/) is introduced, where
Tg − Tf
η=
Tg − Tc
Eq. 7.1
The η values vary between 0 and with as the best film cooling effectiveness. Therefore, the heat flux
ratio can be written as:

q" h Tf − Tw h Tg − Tc
=( )[ ] = ( ) × {1 − η [ ]}
q"0 h0 Tg − Tw h0 Tg − Tw
Eq. 7.2
To obtain any benefit from film cooling, the heat load ratio, q"/q’, should be below 1.0. The heat
transfer coefficient ratio (h/ho) is enhanced due to turbulent mixing of the jets with the mainstream
and is normally greater than 1.0. The temperature ratio (Tf - Tw)/(Tg - Tw), which is related to the film
effectiveness should be much lower than 1.0 such that the heat load ratio is lower than 1.0. The best
169

film cooling design is to reduce the heat load ratio (i.e., smaller h/ho enhancement with greater η) for
a minimum amount of coolant available for a film cooled airfoil 192.
7.1.1.1 End-Wall Film-Cooling
One method of combating the high heat transfer coefficients along the end-wall is through the use of
film-cooling holes whereby cooler air is injected through discrete holes in the end-wall. Film-cooling
hole placement, particularly in the end-wall region, has traditionally been based upon designer
experience whereby a number of design variables are considered. For example, one should take into
account leakage from a slot at the combustor turbine interface whereby cooler gases leak into the
main gas path (combustor film-cooling carryover). Most turbines are designed such that pressures
outside the main gas path are higher than those found
in the main gas path to insure that the hot
combustion gases are not ingested below the
platform. If designed properly, this leakage flow
could be relied upon as a source of coolant. It is also
important to remember that the secondary flows that
develop are affected by the leakage flows at the vane
inlet, as indicated by the previous discussion in this
section.
Other design variables include roughness effects,
film-cooling migration (as will be discussed), film-
cooling limitations resulting from internal cooling
schemes, and cooling hole manufacturing.
Manufacturing of cooling is generally done through
the use of electro-discharge machining (EDM) or
laser drilling with both being subjected to access
limitations. Generally, EDM is more expensive than
laser drilling and conversely laser drilling can tend to Figure 7.4 Measured Adiabatic Wall
be more limited in terms of hole geometries that can Temperatures for Coolant Exiting a
be manufactured. Consider a leakage slot flow Combustor/Vane Leakage Slot (reproduced
between the combustor and turbine whereby the with permission from ASME)
coolant is 0.75% of the core flow. The cooling to the
end-wall that can be provided is shown in Figure 7.4. Coolant exits across much of the width of the
slot albeit in a very non-uniform manner. This non-uniform slot flow arises from the static pressure
distribution and secondary flow development along the end-wall. Although there is a large uncooled
region, often referred to as a bow wake, around much of the vane at the leading edge and along the
pressure side of the airfoil, there is also a well-cooled region in the center of the passage. As the slot
flow increases, the cooling potential also increases to a point after which the benefit is small193.
As was previously mentioned, end-wall film-cooling has largely been based on designer experience.
One difficulty in designing the film-cooling hole pattern is knowing before-hand the local static
pressure along the end-wall, which varies greatly along the end-wall as the flow accelerates through
the passage. If one considers that a single supply feeds all the film-cooling holes and inviscid flow
through the holes, it can be shown that a global, ideal (loss-free) blowing ratio for all the film-cooling
holes is given by the following equation,

192 Je-ChinHan and Srinath Ekkad, “Recent Development in Turbine Blade Film Cooling”, International Journal of
Rotating Machinery, 2001, Vol. 7, No. 1, pp. 21-40.
193 Karen Thole, “Mechanical Engineering Department Penn State University”, Park, PA 16802-1412.
170

ρc Uc ρc Po,c - ps,in
Mglobal = √ .
ρin Uin ρin Po,in - ps,in
Eq. 7.3
where Po,in and ps,in are defined at the inlet to the turbine. If one then wants to compute the local, ideal
blowing ratio for each cooling

ρc Uc ρc Po,c - ps,∞
M= =√ .
ρin Uin ρin Po,in - ps,∞
Eq. 7.4
hole (now taking into consideration that the local velocity varies), the following equation can be used
where ps,∞ is the local static pressure defined at the exit of the cooling hole. Eq. 7.4 indicates that the
same blowing ratio will occur for each hole placed along a constant static pressure line. Designing an
end-wall cooling pattern to achieve a uniform blowing ratio is one methodology that some companies
have used. Figure 7.5 (a) shows a cooling hole pattern whereby the holes in the passage were placed
along a constant static pressure line. Figure 7.5 (b) shows an end-wall cooling hole pattern with
holes that were placed on lines parallel to the incoming flow direction. The end-wall in Figure 7.5
(b) also provides the space for including a mid-passage gap that occurs between vanes as the vanes
are mated on the turbine disk. Vanes are generally cast in either doublets or singlets and then placed
in a turbine disk with some type of seal between the vane platforms. Generally, relatively low levels
of coolant leak from this mid-passage gap (less than 0.3%). Figure 7.5 (c) shows the same film-
cooling hole pattern as that in Figure 7.5 (b) with the exception being the mid-passage gap is
present in Figure 7.5 (c). The coolant flow conditions for all three vanes shown in Figure 7.5 are
the same with 0.75% of the core flow exiting in the form of coolant from the upstream slot, 0.5%

Figure 7.5 Contours of Adiabatic Effectiveness for Two Film-Cooling Hole Patterns (left and center)
With a Mid-Passage Gutter for the Cooling Hole Pattern in the Center (Right) (Reproduced With
Permission From the Publisher of ASME)
171

coolant exiting the film-cooling holes and no net coolant flow through the mid-passage gap for Figure
7.5(c). The contours plotted in Figure 7.5 represent the non-dimensional adiabatic wall
temperatures. As discussed in the previous section, adiabatic surface temperatures represent the
local fluid temperature. From all three contour plots in Figure 7.5, it is clear that the upstream slot
can definitely be an important part of cooling mid portion of the end-wall. In comparing Figure 7.5
(c) to the rest, however, it is also seen that the mid-passage gap limits the area of the cooling present.
The mid-passage gap provides a convective coolant trough through which coolant from the upstream
slot enters and passes through the gap along the mid-passage before exiting the gap near the trailing
edge of the airfoil.
The streamlines superimposed upon the experimental measurements of the adiabatic wall
temperatures presented were computed using CFD. These streamlines were extracted from the CFD
simulations at a location very near to the end-wall (less than 2% of the span). The streamlines, for
the most part, can be used to predict the trajectory of the coolant flow exiting the cooling holes. There
are some regions, however, where the streamlines differ from the jet trajectories, such as at the
entrance to the passage closer to the pressure side. In comparing all of the cooling hole patterns in
Figure 7.5, it is clear that more uniform cooling can be achieved by placing cooling holes along
constant static pressure lines as in Figure 7.5 (a). All three patterns illustrate the difficulty of cooling
the end-wall along the pressure side of the vane and along the leading edge-end-wall juncture. In
general these are very difficult areas to cool because of the secondary flows that were described
previously. The horseshoe vortex in the stagnation region can lead to coolant being swept upstream
of the holes at low blowing ratios, whereas at high blowing ratios the coolant separates from the end-
wall and impacts the vane surface rather than the end-wall surface. Despite the cooling holes injecting
coolant towards pressure side of the vane, the sweeping motion of the passage vortex prevents
cooling at the juncture between the pressure side of the vane and the end-wall194.
As an animation example on the right, courtesy of siemens blog,
consider the cooling flow which is meant to blanket and protect the
blade surface. The turbulence between the primary flow and the
blanket of cooling flow streams has a massive impact on the GTFilmCooling.mp4
effectiveness in shielding the metal parts from high heat loads.

7.2 Blade Cooling Using Vanes in Blades


Depending on the type of machine,
physical and geometrical effects have to
be taken into account. A complicating
factor is that it is necessary to carry out
parametric studies considering several
geometric options in the process of
designing the cooling systems. This
normally takes a lot of time to generate
mesh models due to the mesh resolution
required in the boundary layer. This task
can be achieved with CAD embedded
technology and automatic meshing to
significantly reduce the overall simulation
time and allow them to frontload their
Figure 7.6 Vane Section with Ten Cooling Channels and
design process, as argued in [Rubekina et, Temperature Distribution Computed

194 Karen Thole, “Mechanical Engineering Department Penn State University”, Park, PA 16802-1412.
172

al.]195. The locations of vanes and their characteristics are most important. More recently there was
an inquiry by [ Jasim et al.]196 where it was mainly focused on gas turbine blade heat transfer analysis
and effect of increasing the number of external film cooling holes rows.
7.2.1 Vane NASA C3X
The first case is the NASA C3X Vane
Experiment. The experimental
object was represented as a linear
array. Each cascade employed three
vanes, characteristic of a first-stage
turbine (see Figure 7.6). Geometry
parameters of vane and cooling
channels were taken from [Hylton et
al. (1983)]. Each of the vanes was
cooled by an array of ten radial
cooling holes. The vanes were
fabricated of ASTM type 310
stainless steel, which has a relatively
low thermal conductivity197. The Figure 7.7 Surface Temperature Distribution on the Suction
average heat transfer coefficient for Side (Left) and The Pressure Side (Right) of the Vane
each cooling hole was calculated
from empirical relation of Nusselt
Number with Reynolds and Prandtl Numbers as:

D
NuD = F(0.022. Pr 0.5 . Re0.8
D ) = α.
λ
Eq. 7.5
Where D is the hole diameter, λ is thermal conductivity. The Re for each cooling channel was
determined from the measured flow rate, the cooling hole diameter, and viscosity based on the
average coolant temperature. F is a function of Pr, ReD, and x/D, which corrects the Nu expression for
a fully developed thermal boundary layer to account for thermal entrance region effects. Each of the
vanes was cooled by an array of ten radial cooling holes.
Another test cases with more vanes is the Convective Heat Transfer in The Steam Turbine Vane as
depicted in Figure 7.7. The investigations on the thermal efficiency potential of steam cooling, have
been carried out experimental work [Bohn et al (2002)]. Geometrical configuration consists of a
rectangular duct containing a three-vane cascade. The central vane can be convectively cooled by
supplying steam to 22 straight radial cooling passages198.

195 A.V. Rubekina, A.V. Ivanov, G.E. Dumnov, A.A. Sobachkin (Mentor Graphics Corp., Russia); K.V. Otryahina
(PAO NPO Saturn, Russia), “Keeping it Cool in Gas Turbines”, Mentor Graphics white paper.
196 Hadeel Raheem Jasim, Narsimhulu Sanke and Khaled Al-Farhany, “Heat Transfer Simulation of Gas Turbine

Blade with Film Cooling”, International Journal of Modern Engineering and Research Technology, 2018.
197 A.V. Rubekina, A.V. Ivanov, G.E. Dumnov, A.A. Sobachkin, “ Using Modern CAD-Embedded CFD Code For

Numerical Simulation Of Heat Transfer In Vanes And Blades”, Mentor Graphics White Paper.
198 A.V. Rubekina, A.V. Ivanov, G.E. Dumnov, A.A. Sobachkin, “ Using Modern CAD-Embedded CFD Code For

Numerical Simulation Of Heat Transfer In Vanes And Blades”, Mentor Graphics White Paper.
173

cooling air temperature solid temperature on the pressure side

Figure 7.8 Flow Streamlines Colored Cooling Air Temperature into Passages

7.2.2 Showered Film Type Cooling


The second case consisted of a showerhead film cooled vane, five rows of staggered cylindrical
cooling holes is used in [Nasir et al (2008)]. The vane cascade consists of four full vanes and two
partial vanes. The number of cooling holes on the stagnation region row is 17. Cooling flow is injected
at 90° angle to the freestream and 45° angle to the span of the vane. Overview and dimensions of the
test vane, as well as, Gas temperature distribution is presented in Figure 7.8 199.
7.2.3 Complex Network of Vanes
The final case is a rotor blade used by NPO Saturn. The detailed description of similar blade can be
found in [Vinogradov et al]. The blade is made of heat-resistant alloy ZS32. Ceramic coating (a
thickness of 0.02-0.03 mm with thermal conductivity k=2-3 W/(m ·K)) serves to insulate components
from large and prolonged heat loads by utilizing thermally insulating materials. Relatively cold air is
passed through the passages inside the turbine blade. Complicated internal cooling system includes
serpentary channels with rib tabulators. Then coolant goes out the blade through holes located on
the blade tip. The remaining coolant is ejected from the trailing edge of the profile. As for previous
test cases evaluation of the temperature distribution of the blade was made by means of conjugate
heat transfer simulation. Ceramic heat barrier coating was taken account as the thermal resistance
with equivalent parameters. Numerical simulation included modeling of mainstream gas flow, air
flow through the internal channels of the blade as well as heat transfer in solid and between fluid and
solid by convection. Shown in Figure 7.8 cooling air temperature distributions along flow
streamlines as well as surface metal temperature field on both side of the blade are in a good
concordance with qualitative evaluates. Recently there was an investigation by [Hasanpour et al.]200
regarding hole configuration effect on turbine blade cooling, where first one of the commonly used
cooling hole geometry is investigated; cylindrical holes and then two other configurations are

199 Explanation of Figure 7.6 - (top left) and section view of stagnation row of holes (top right), as well as Gas
temperature distribution (bottom left), cooling air streamlines with the Mach number distribution (bottom
right)
200 A.Hasanpour, M. Farhadi and H.R. Ashorynejad, “Hole Configuration Effect on Turbine Blade Cooling”, World

Academy of Science, Engineering and Technology 49 2011.


174

simulated. The average temperature magnitude in mid-plan section of each configuration is obtained
and finally the lower temperature value is selected such as best arrangement.

7.3 Conjugate Heat Transfer


In CHT problems, the URANS equations are solved at all internal fluid cells, whereas the heat
conduction equation is solved at all internal solid cells using the same spatial discretization and time
marching numerical algorithm; the interface boundary conditions require that both the temperature
and heat-flux are the same for the fluid and the solid at all boundary points:

Tw,f = Tw,s , k sTs .n w = k f Tf .n w Eq. 7.6

where the subscripts s and f refer to solid and fluid respectively, k indicates the thermal conductivity,
and nw is the unit vector normal to the wall. In order to impose the boundary conditions above, at the
solid interface cells the effect of the boundary conditions imposed by the flow upon the temperature
and heat-flux at the body surface has to be accounted for, exactly as the same conditions are to be
enforced at the fluid-interface cells. A direct coupled solution of the URANS and heat conduction
equations is unfeasible. Thus, an iterative procedure is to be employed. A first-order-accurate
approximation is obtained as:

Tf .n w,f = (Tint,f - Tw,f )  f


m m -1

Eq. 7.7
where m is the current iteration and βf is the inverse of the distance between the fluid interface cell
and the wall Eq. 7.7 provides the heat fluxes (apart from the thermal conductivity) at all fluid cells.
Such values are then used to provide those pertaining to all Solid Cell Projection Point (SCPP) by
means of a distance-weighted interpolation and then, the Neumann condition at all SCPPs201.
7.3.1 Case Study - Heat Transfer in Separated Flows on the P. S. of Turbine Blades
Heat transfer in separated flows on the pressure side of a typical high lift turbine profile is
numerically investigated by [De La Calzada, et, al.] . The numerical code was first validated on
attached flows in turbine blades. To obtain flow separation cases, the profile is subject to large
negative incidences so that a separation bubble is obtained at the pressure side. The numerical
results are compared to available experimental data for code validation. The aim of the present
investigation is to perform a detailed numerical study of the heat transfer phenomena in separated
flows at flow conditions representative of LPT airfoils. A comparison with experimental data is
performed, hence allowing the validation of the code and the confirmation of the main flow features.
The relationship between the dynamic and thermal boundary layers and their importance of the
velocity component perpendicular to the wall in creating injection of flow towards the wall or
ejection of flow from the wall and their effect in the heat transfer is analyzed.
7.3.1.1 Literature Survey
Much attention has been paid to the investigation of large flow separation in simple cases, including
both velocity related measurements and heat transfer measurements. These include experimental
investigations of backward-facing steps as [Vogel and Eaton]202 or [Sparrow et, al.]203, where the

201 D. De Marinisa, M. D. de Tullioa, M. Napolitanoa, and G. Pascazio, “A conjugate-heat-transfer immersed-


boundary method for turbine cooling”, Energy Procedia 82 ( 2015 ) 215 – 221.
202J. C. Vogel and J. K. Eaton, “Combined Heat Transfer and Fluid Dynamic Measurements Downstream of a

Backward-Facing Step “, Heat and Mass Transfer, vol. 107, pp. 922-929, 1985.
203 E. M. Sparrow, S. S. Kang, and W. Chuck, Relation Between the Points of Flow Reattachment and Maximum

Heat Transfer for Regions of Flow Separation, Int. J. Heat Mass Transfer, vol. 30, no. 7, pp. 1237-1246, 1987.
175

relationship between the separation region and the heat transfer features was studied.
Corresponding numerical investigations have been performed on similar configurations like the one
by [Kaminejad et al.]204 where only laminar conditions and very low Reynolds numbers are
considered. The effect of turbulence was taken into account for example by [Rhee and Sung]205, where
good agreement with experimental data was also found for very low Reynolds numbers. More
recently, Rhee and Sung also investigated the effect of local forcing on the separation and reattaching
flow. However, very little attention has been paid to the heat transfer in large separated flow regions
in turbine representative conditions. [Bassi et al.]206 present CFD results on the separated flow region
of a HPT airfoil with cutter trailing edge with no cooling ejection, but only a short discussion about
the separated flow physics is included. Regarding experimental investigations, [Rivir et al.]207 have
measured the flat plate heat transfer in a region of turbulent separation, and [Bellows and Mayle]208
have measured the heat transfer on a blunt body leading edge separation bubble both for cases of
high Reynolds number209. More recently, [Calzada and Alonso]210 performed a numerical
investigation of large flow separation region at the pressure side of a turbine profile but not
comparison with experimental results was included. [Lutum and Cottier] presented a similar
investigation, but results indicated that simulations were not able to reproduce experimental heat
transfer results at the pressure side separation region especially for low turbulence levels. From
experimental investigations, it is known qualitatively that separated flow regions are characterized
by large and rapid variations of the heat transfer (e.g., Rhee and Sung). Furthermore, the heat transfer
presents a local minimum and a local maximum in the vicinity of separation and reattachment points
respectively, with regions where the heat transfer coefficient (HTC) is much larger than that of
attached flows211. Taking into account that separated flow regions are usually characterized by high
turbulence levels and large scale unsteadiness, there is a tendency in the heat transfer community to
explain the heat transfer phenomena in separated flows in terms of the generation of turbulence
rather than in terms of the dynamic and thermal boundary layers relationship212.
7.3.1.2 CFD Modeling
A CFD solver used based on [Jameson et al]213 been used. Convective terms are discretized using a
cell vertex scheme, and the viscous terms are computed by means of the Hessian matrix. Integration
in time is performed using an explicit five stage Runge-Kutta scheme where the viscous and artificial
dissipation terms are evaluated in the first, third, and fifth stages. The code runs on unstructured

204 H. Kazeminejad, M. Ghamari, and M. A. Yaghoubi, “A Numerical Study of Convective Heat Transfer from a
Blunt Plate at Low Reynolds Number”, Int. J. Heat Mass Transfer, vol. 39, no. 1, pp. 125-133, 1996.
205 G. H. Rhee, and H. J. Sung, “A Low-Reynolds Number, Four Equation, Heat Transfer Model for Turbulent

Separated and Reattaching Flows”, Int. J. Heat Fluid Flow, vol. 18, pp. 38-44, 1997.
206 F. Bassi, S. Rebay, M. Savini, S. Colantuoni, and G. Santoriello, “A Navier-Stokes Solver with Different

Turbulence Models Applied to Film-Cooled Turbine Cascades”, Paper No. 41, AGARD-CP-527, 1993.
207 R. B. Rivir, J. P. Johnston, and J. K. Eaton, “Heat Transfer on a Flat Surface under a Region of Turbulent

Separation”, Turbomachinery, vol. 116, pp 57-62, 1997.


208 R.J. Bellows and R. E. Mayle, “Heat Transfer Downstream of a Leading Edge Separation Bubble,”,

Turbomachinery, vol. 108, pp. 131-136, 1986.


209 W. Merzkirch R. H. Page, and L. S. Fletcher, “A Survey of Heat Transfer in Compressible Separated and

Reattached Flows”, AIAA Journal, Vol. 26, no. 2, pp. 144-150, 1988.
210 P. De La Calzada and A. Alonso, “Numerical Investigation of Heat Transfer in Turbine Cascades with Separated

Flows”, Turbomachinery, vol. 125, no. 2, pp. 260-266, 2003.


211 R. B. Rivir, J. P. Johnston, and J. K. Eaton, “Heat Transfer on a Flat Surface under a Region of Turbulent

Separation”. Turbomachinery, vol. 116, pp 57-62, 1997.


212 P. De La Calzada, M. Valdes, and M. A. Burgos, “Heat Transfer in Separated Flows on the Pressure Side of

Turbine Blades”, Industria de Turbopropulsores S. A., Spain.2007.


213 A. Jameson, W. Schmidt, and E. Turkel, “Numerical Solution of the Euler Equations by Finite Volume Method

using Runge-Kutta Time Stepping Schemes”, AIAA, Paper 81-1259, 1981.


176

meshes which are built by a quasi-structured layer all along the walls, where viscous effects are
expected to be dominant and by a fully triangular unstructured mesh in the rest of the flow domain
obtained by Steiner triangulation214. For turbulence simulation, the two equations κ-ω model from
Wilcox215 is implemented. More details of the numerical code can be found in Corral and Contreras216.
Numerical results are post processed to obtain heat and mass transfer relevant parameters at the
walls. This is performed by computing velocity and temperature variation in the direction
perpendicular to the wall so that heat transfer and friction coefficient as well as Stanton number can
be computed, as defined below. The local Stanton number is the equivalent for the temperature to
the local skin-friction coefficient for the velocity. Although the local Stanton number variations do
not represent variations in heat flux alone but also take into account the local velocity value, it is the
most adequate parameter to describe the thermal boundary layer behavior and to develop special
correlations for heat transfer estimation (i.e., Reynolds-Colburn analogy).

 T 
− k 
HTC x  n  w τw
Ch x = = , Cf x = Eq. 7.8
ρ e C p u e (T0 − Tw )ρ e C p u e 1
ρ e u e2
2
Note, that the total temperature is used in the definitions instead of the adiabatic wall temperature,
even though compressible effects and therefore viscous dissipation may be important since the
representative cases for LPT usually imply an exit Mach number of around 0.5, as we have in our
study. However, the difference between the aforementioned coefficients and the corresponding
compressible definitions can be kept sufficiently low if the wall temperature for the computations is
properly chosen. In our particular cases, the total temperature is defined as in the experiments and
the wall temperature is taken around 25 K higher than the fluid temperature, which keeps the
difference between compressible and incompressible heat and mass transfer coefficient values lower
than 2% even at regions with Mach numbers around 0.5. This wall temperature value also develops
a thermal boundary layer whose magnitude is large enough to avoid high sensitivity to any random
numerical errors in the resolution of the temperature field around the wall.
7.3.1.3 Description of the Blade Computational Grids and Results for Attached Flow
The T106-300 blade section has been used as a generic geometry representative of a typical highly
loaded LPT airfoil217 (see Figure 7.9). In this investigation, the blade profile is subject to extremely
large negative incidences in order to have a large separation bubble on the pressure side. Mach and
Reynolds numbers are varied around typical LPT values. The generated grid is hybrid in nature with
higher definition in regions adjacent to the wall, trailing edge and leading edge, as shown in Figure
7.9. Due to the expected flow separation at the pressure side when the profile is subject to high
negative incidence, on this investigation the viscous mesh is extended to a region larger than attached
flows would require for this Reynolds number. The objective of this large region with high definition
is to capture the shear layers and flow features on the pressure side large bubble. However, in order

214 R. Corral and J. Fernandez-Castañeda, “Surface Mesh Generation by Means of Steiner Triangulation”, Proc.
29th AIAA Fluid Dynamics Conference, vol. 39, pp. 176-180, Albuquerque, New Mexico, 1998.
215 D. C. Wilcox, “Reassessment of the Scale Determining Equation for Advanced Turbulence Models”, AIAA J.,

vol. 26, pp. 1299-1310, 1988.


216 R. Corral and J. Contreras, “Quantitative Influence of the Steady Non-Reflecting Boundary Conditions on Blade-

to-Blade Computations”, Proc. 45th ASME Gas Turbine and Aero engine Congress, Exposition and Users
Symposium, ASME Paper 2000-GT-515, Munich, 2000.
217 H. Hoheisel, “Test Case E/CA-6, Subsonic Turbine Cascade T106, Test Cases for Computation of Internal Flows

in Aero Engine Components”, AGARD-AR-275, 1990.


177

to avoid any mesh sensitivity the same grid


consisting of 8,623 nodes was kept unchanged for
all cases, including the attached flow achieving a
range of y+ values at the pressure side in the order
of y+< 3. Results at design conditions are shown
pressure distribution is considered to match well
with the experiments, in particular on the major
part of the suction side.
Since we are interested mainly in the flow along
the pressure side, no attention will be paid to the
separation bubble at the back suction surface that
the code does not predict probably due to a too
soon turbulence generation and boundary layer
transition. Although the level of pressure
achieved by the numerical results at the pressure
side of the profile is lower than the experimental
data, the heat transfer level matches well with the Figure 7.9 2D Hybrid Mesh around the T106
experiments. However, more HTC oscillations are Blade
found in the experiments compared with the
smoother results predicted by the numerical simulation. It is interesting to note that the heat transfer
measurements at the acceleration region of the leading edge decrease to lower values than the CFD
results. This might indicate that the profile is subject to a slightly larger negative incidence in the
experiments, hence creating an acceleration-deceleration behavior achieving a higher final pressure
as shown by the results. The final acceleration region towards the trailing edge has a more
pronounced effect on the numerical simulation, where the heat transfer value shows higher increase
due to the expected thinning of the boundary layer with the increase in the external velocity.
7.3.1.4 Separated Flow with Large Separation Bubble
Results at the extreme conditions of 37.7° negative incidence (i.e., inlet angle β1 = 90°) are shown
first. The comparison between numerical and experimental results in terms of pressure distribution

Figure 7.10 Blade Profile vs. Pressure Coefficient (Courtesy of De La Calzada et al.)
178

along the airfoil is presented in Figure 7.10. The separation region is characterized by low velocities
and a fairly constant pressure distribution. However, at the reattachment region the static pressure
increases reaching a local maximum (i.e., local minimum value of Cp in Figure 7.10), which indeed
indicates the reattachment point. The experimental results indicate, to some extent, a shorter
separation bubble which reattaches earlier, hence starting earlier also the acceleration towards the
trailing edge. This may be related to the already identified higher pressure at the pressure side
predicted by the CFD, which might indicate some slight difference in local incidence angle between
the experiments and the simulations 218.
Some more detail about the flow can be identified by comparing experimental and numerical heat
transfer results shown in Figure 7.10. Both experimental and numerical results show two local
minima and maxima between the extreme values achieved at leading and trailing edges (these

(B) Pressure Contours

(A) Mach No.

(C) Temperature (D) Velocity vectors

Figure 7.11 Flow Field at the Front and Middle Parts of the Separation Bubble (Courtesy of De La
Calzada et al.)

218P. De La Calzada, M. Valdes, and M. A. Burgos, “Heat Transfer in Separated Flows on the Pressure Side of
Turbine Blades”, Industria de Turbopropulsores S. A., Spain, 2007.
179

extreme values are not shown in the graph). The first minimum occurring at around 0.03 x/L
corresponds to the expected reduction in heat transfer rate. In Figure 7.11 (A-D) the Mach number,
total Pressure, Temperature, Velocity vectors fields are plotted where the large bubble at the
pressure side can be clearly identified. Helped by the streamlines traces, the multiple bubble
configuration can be also identified. In this particular case, two bubbles appear. As confirmed in
Figure 7.11(D) by the velocity vectors, one small bubble is stretched towards the pressure wall,
developing at the center of the full separation region whose vortex is rotating counterclockwise, and
one large bubble, rotating clockwise is extending up to the external shear layer along the major part
of the pressure side, having its vortex core at the rear part of the separation region while extending
its vortex influence also to the front part. Detail of the temperature field and flow velocities in the
regions of flow separation and reattachment are plotted in Figure 7.11 (C-D). At points 1 and 3, the
flow is separating and a large component of the velocity perpendicular and directed away from the
wall exists. This flow configuration takes heat from the side walls and ejects it, creating an ejection
stagnation or fountain-like region where the wall thermal field is penetrating the main flow helped
by the perpendicular component of the velocity, hence increasing the effective thermal boundary
layer and decreasing the heat transfer rate.
This phenomena is particularly clear in front of point 3, where the increase in the thermal boundary
layer thickness can be easily identified by the extension of the high temperature region close to the
wall in Figure 7.11. Point 3 corresponds to the separation of an internal second bubble, which must
also exist in the experiment configuration since the local minimum can be also identified in
measurements in Figure 7.11. Points 2 and 4 correspond to reattachment points where there is an
important component of the velocity perpendicular and towards the wall, hence taking fresh fluid to
the wall and creating an injection stagnation region where the thermal boundary layer is reduced
and heat transfer is increased. To further investigate the thermal boundary layer developing through
the separation region the temperature profiles developing along straight lines perpendicular to the
wall, marked a-d in Figure 7.11 (B). Note, that the y coordinate is non-dimensionalized with the
thermal boundary layer thickness, which is basically coincident with the thickness of the separated
region. Dotted lines represent the temperature distribution of the corresponding adiabatic wall case
where static temperature only varies as a result of the velocity profile (being the stagnation
temperature fundamentally constant), while solid lines show the temperature distribution within the
re-circulation region at different distances from the wall in the case with heat transfer and heated
wall219.
The lowest wall temperature gradient is obtained at the leading edge separation point 1 (line a),
where even with the thermal boundary layer being relatively thin the fluid temperature shows a low
gradient specially close to the wall driven by the ejection of heated flow from the wall through the
ejection stagnation region configuration. On the contrary, the highest heat flux is achieved at the main
bubble reattachment point 4 (line d) where, additionally to the thin thermal boundary layer
thickness, the fluid temperature variation is mainly concentrated at the wall in a region about 10%
of that thickness, hence increasing the temperature gradient at the wall. This reduction of the
effective boundary layer thickness is driven by the injection stagnation region configuration, where
the velocity component perpendicular to the wall is forcing the thermal boundary layer to be
squeezed towards the wall. As an additional proof showing that there is low coupling between the
dynamic and thermal boundary layers and their gradients in separated regions, the relationship
between the velocity parallel to the wall and temperature gradients at the wall is investigated. It is
widely accepted that the Reynolds-Colburn analogy is only reliable in attached flows only for modest,
near-zero, pressure gradients, and with a constant wall temperature. The computed local skin-

219P. De La Calzada, M. Valdes, and M. A. Burgos, “Heat Transfer in Separated Flows on the Pressure Side of
Turbine Blades”, Industria de Turbopropulsores S. A.,.Spain, 2007.
180

Figure 7.12 Heat Transfer Coefficient for Different Negative Incidences (Courtesy of Calzada et al.)

friction coefficient (absolute value), the Stanton number, and the Reynolds-Colburn analogy are
shown in Figure 7.12 to further demonstrate that the analogy between dynamic and thermal
boundary layer is not valid for separated flows even when no pressure gradient exists. Only at the
rear acceleration region where attached flow is ensured, the Reynolds analogy tends to follow the
correlation showing a conventional relationship between dynamic and thermal boundary layers and
their gradients at the wall. Furthermore, unlike the Stanton number, the skin friction approaches
zero not only at the separation and reattachment regions but also along the major part of the
separated flow region, hence confirming that the Reynolds-Colburn analogy is not applicable. This is
clear proof that there is a very weak coupling between velocity parallel to the wall and thermal
boundary layers in separated flows. On the contrary, it is the convective transport of fluid in a
direction normal to the wall and the fluid conduction effects in low velocity regions what drive the
heat transfer phenomenon, hence supporting once again the prime role of the stagnation region
configurations on the heat transfer mechanism.
7.3.1.5 Inlet Flow Angle Effects
Inlet flow angles of 90°, 100°, 110° (i.e., -37.7, -27.7°, and -10.7° incidence angle, all with separated
flows at pressure side), and 127.7° (i.e., 0° incidence angle, with pressure side attached flow) have
been simulated for the nominal isentropic exit Re = 1.5 x 105 and an isentropic exit Mach number of
0.5. The results in terms of HTC and Stanton number are presented in Figure 7.12 and Figure
7.13220. As expected, the size of the bubble is decreasing with the reduction of the negative incidence
angle as can be concluded from the location of the maximum values of Stanton numbers in Figure
7.13. All the I separated flow cases show relatively large bubbles varying the reattachment points
from 0.5 x/L for -17.7° incidence up to 0.6 x/L for -37.7° incidence. Unlike the HTC whose local
maximum value at reattachment point is maintained almost constant along the attached acceleration
region up to the trailing edge region, the Stanton number clearly generates a more pronounced local
maximum value at the reattachment point driven by the combination of maximum heat flux and static
pressure (i.e., minimum external velocity).

P. De La Calzada, M. Valdes, and M. A. Burgos, “Heat Transfer in Separated Flows on the Pressure Side of
220 220

Turbine Blades”, Industria de Turbopropulsores S. A., Spain, 2007.


181

It is noticeable that in all separated flow cases the same multiple bubble configuration is obtained as
indicated by the presence, within the separated region, of one additional local minimum and one
additional local maximum in heat transfer parameters although it is less evident in the slight negative
incidence case (i.e., -17.7° incidence). Generally, it can be concluded that the separated region always
generates a redistribution in the heat flux by decreasing the value at the front separation region and
by increasing it at the rear reattachment region. The higher or lower surface averaged effective value
will depend on the particular geometry and conditions.

Figure 7.13 Stanton Number for Different Negative Incidences (Courtesy of De La Calzada et al.)

7.3.1.5.1 Reynolds Number Effect


Reynolds number effect is investigated by simulating different cases with the same incidence and exit
Mach numbers but different fluid conditions so that the Reynolds number is changed. In order to
achieve the required effect in the simulations, only the pressure level is modified. Figure 7.15 and
Figure 7.14 present the results when the Reynolds number is varied between 150,000 and 400,000.
The dependence of heat flux on Reynolds number at separating and impingement regions can be
obtained analytically on simple cases. At impingement points the case of plane and axisymmetric
laminar flows can be integrated to obtain the known dependence of the heat flux on the Reynolds
number to the power of 0.5221. Similarly, at separating points expansion equations can be obtained
which also show a dependence on Reynolds number to the power of 0.5 in simple cases as wedge and
Howarth's decelerating flow 222. However, these methods are of difficult application to complex cases
as presented here where a shear flow impinges on the wall with an inclination angle (i.e., the bubble
reattachment occurring in our case), or where the separation occurs within a region of already
separation bubble (i.e., the secondary bubble appearing in our case).
In our simulation cases, the HTC (i.e., heat flux) also increases with Reynolds number as expected.
However, the maximum values at the reattachment point in these cases increases with an exponent
approximately equal to 0.3. It is interesting to note that the Stanton number varies inversely with the

F. M. White, “Viscous Fluid Flow”, McGraw-Hill, pp. 162 and 248, 1991.
221

H. W. Kim and D. R. Jeng, “Convective Heat Transfer in Laminar Boundary Layer Near the Separation Point”,
222

ASME Proc. of the 1988 National Heat Transfer Conference, vol. 3, pp. 471-476, 1988.
182

Figure 7.15 Stanton Number vs. Reynolds Numbers (Courtesy of De La Calzada et al.)

Reynolds number, as shown in Figure 7.14. By applying the definition relationship between HTC
and Stanton number Ch, it can be seen that the ratio HTC/Ch must retain a dependence on Reynolds
to the power of 1. Therefore, the Stanton number dependence on the Reynolds number should vary
with and exponent of -0.7 in these cases according to the exponent 0.3 found for the HTC, which is
indeed confirmed by the maximum values at the reattachment point shown in Figure 7.14.

Figure 7.14 Heat Transfer Coefficient vs. Reynolds Number (Courtesy of Calzada et al.)
183

One interesting feature is that, for all Reynolds numbers investigated, the size of the bubble and the
internal structure (i.e., multiple bubble configuration) is the same, as can be concluded from the
location of local maxima and minima in the figures. Although it could be expected that increasing the
Reynolds number would reduce the size of the separation bubble, in this particular case the
reattachment is driven by the acceleration of the flow and the role played by the Reynolds and the
corresponding boundary layer instability and potential transition is expected to be very minor. This
is a completely different behavior compared to cases in which there is no flow acceleration, and the
reattachment is driven by boundary layer transition.
In these latter cases, the increase of turbulence produces an early transition and reattachment and
an increase in heat flux due to the stronger reattachment vortex on a blunt flat plate subject to
pulsating conditions. Although the implemented numerical turbulence model was able to produce
high turbulence and the corresponding boundary layer transition on the suction surface to avoid the
back surface separation, at the pressure side the turbulence generation is concentrated on the
external shear layer and from there it is convected downstream to the trailing edge and the
downstream wake. Therefore, it is thought that in cases like this the turbulence is not a strong enough
mechanism to force sufficient flow entrainment and perturbation to the shear layer to produce an
early reattachment of the boundary layer, and it is then expected that the size of the separation
bubble will depend weakly on the Reynolds number and turbulence. See Figure 7.15.
7.3.1.6 Concluding Remarks
A better understanding of the flow physics and the heat transfer mechanisms in large separated flow
regions have been achieved by means of a numerical investigation on the T106-300 typical LPT airfoil
subject to large negative incidence. Flow separation is characterized by a pronounced reduction in
HTC at the separation region, close to the leading edge where the minimum value is achieved, and by
an increase at the reattachment region where the maximum value is achieved. Those are extreme
values, much lower and higher than the ones obtained for attached flows. It is concluded that the
velocity component perpendicular to the wall is the main contributor to the generation of ejection
and impingement stagnation configurations, where the flow is taken from or towards the wall, hence
affecting the thermal field in those regions and contributing to create a lower or higher temperature
gradient at the wall and the corresponding HTC values.
By analyzing the Reynolds-Colburn analogy all along the pressure side of the profile, the low coupling
between the velocity component parallel to the wall and the thermal field and their gradients within
the separation region is confirmed. Additionally, it is also shown that an important variation in HTC
values can occur within the separation region due to the presence of secondary separation bubbles
which can create additional separation and reattachment points. This is confirmed by both numerical
and experimental results for the high negative incidence (i.e.,-37.7° incidence), which show the
presence of one additional local maximum and one local minimum in HTC values that must indicate
the presence of additional corresponding reattachment and separation points, hence indicating the
presence of and the additional secondary separation bubble. Moreover, it is also shown by the
numerical results that the multiple bubble configuration is found for all separated cases investigated
here (i.e., negative incidence varying from -17.7° to -37.7°). Finally, the variation of the heat transfer
with the Reynolds number is investigated. The numerical results show no variation of the separation
bubble size with Reynolds number varying from 150,000 to 400,000. A dependence of the HTC on
the Reynolds number to the power of 0.3 is obtained in the separation region, in particular at the
maximum value occurring at the main bubble reattachment point on the rear part of the separation
bubble223.

223P. De La Calzada, M. Valdes, and M. A. Burgos, “Heat Transfer in Separated Flows on the Pressure Side of
Turbine Blades”, Industria de Turbopropulsores S. A., Madrid, Spain.
184
185

8 General Perspectives on Turbulence Consideration


Recent advancements in turbulence modeling have brought CFD improvements in turbulence
improvements to the forefront of turbo-machinery design and analysis. Today, 3D Euler, Quasi-3D
viscous, and 3D full Navier-Stokes analyses are integral parts of turbo-machinery design. Fan rotors
are designed using viscous 3D CFD models using these models the blade geometry is tailored to
control shock location, boundary layer growth, and end-wall blockage. The complexity of the turbo-
machinery flow field, so far, limits CFD simulations to Reynolds averaged (RANS) approximations.
The flow field of a transonic fan over its entire operating range is particularly troublesome; it contains
all the flow aspects most difficult to represent; boundary layer transition, separation, shock-
boundary layer interactions, and large flow unsteadiness. Multistage configurations extends the
complexity as “neither in the stator nor rotor frame of reference are steady in time”. Direct Numerical
Simulations (DNS) and Large-Eddy Simulations (LES) are not currently practical for the
fan/compressor flow field. The DNS explicitly solves for the instantaneous flow field and requires
extremely fine gridding to resolve the smallest length scales; on the order of Re9/4. Thus, state-of-the-
art turbo-machinery CFD involves solution of the RANS equations and hence, some modeling of the
physics. The Reynolds-averaging process-the decomposition of the instantaneous flow field into
mean and fluctuating components and subsequent temporal averaging introduces more unknowns
than available equations for solution. Key modeling aspects are associated with this so-called
“turbulent closure problem.” To obtain mathematical closure, the Reynolds stress terms must be
related to mean flow properties either empirically or through a flow model which allows calculation
of this relationship (eddy “viscosity,” mixing length, transport equations). The large majority of
computational schemes for turbomachinery currently involve the use of the linear (Boussinesq)
relationships between stresses and strains,

∂ui ∂uj 2
−ρu′i u′j = μT ( + ) − ρκδij
∂xj ∂xi 3
Eq. 8.1
coupled with algebraic expressions or, at most, differential equations for the turbulent velocity and
length scales to which the turbulent viscosity is related. This framework is accepted as being
adequate for thin shear flows and is able to reproduce transition in simple boundary layers, if
combined with appropriately constructed and calibrated transport equations for the variation of the
scales in low-Reynolds-number conditions. However, it fails to resolve turbulence anisotropy and to
represent correctly the effects of normal straining and curvature on the turbulent stresses.
As an example of 1st order turbulence models, [Simões, et. al.]224 compare different turbulence
models available in commercial codes such as κ-ε, κ-ω and Mentors SST. It was concluded that the
SST model was the most accurate results among the three models evaluated, when comparing them
to the experimental data of the compressor performance as depicted in Figure 8.1. The principal
aerodynamic characteristics of most turbomachine flows are governed mainly by a balance between
pressure gradient and convection, while turbulence here tends to affect mainly secondary flow
features and the losses225. This is at least so in low-load conditions in which the boundary layers are
relatively thin and attached. In high-load and off design conditions, however, turbulence can
contribute substantially to the aerodynamic balance and is thus a process of major practical interest.
In such circumstances, the boundary layers grow rapidly, separation can ensue on both suction and

224 Marcelo R. Simões, Bruno G. Montojos, Newton R. Moura, and Jian Su,” Validation of Turbulence Models For
Simulation of Axial Flow Compressor”, Proceedings of COBEM 2009.
225 W.L. Chen, F.S. Lien, M.A. Leschziner, “Computational prediction of flow around highly loaded compressor

cascade blades with non-linear eddy-viscosity models”, International Journal of Heat and Fluid Flow 19, 1998.
186

pressure sides (depending on the blade geometry and the incidence angle) and stream wise vorticity
is intense. The sensitivity of major mean flow features to turbulence is especially high when the flow
enters the blade passage at an angle which departs materially from the design value, thus causing
leading-edge separation and high flow displacement, followed by transition in the separated shear
layer. More generally, transition tends to be a highly influential process in the majority of off-design
flows in that details of the location and evolution of transition can dictate the sensitive response of
the boundary layers to pressure gradients.

Figure 8.1 Pressure Ratio by Normalized Mass Flow (Courtesy of Simoes)

This framework is accepted as being adequate for thin shear flows and is able to reproduce transition
in simple boundary layers, if combined with appropriately constructed and calibrated transport
equations for the variation of the scales in low-Reynolds-number conditions. However, it fails to
resolve turbulence anisotropy and to represent correctly the effects of normal straining and
curvature on the turbulent stresses. The last two deficiencies are especially important in blade flows;
first, because the state of turbulence at the leading-edge impingement region is crucially important
to the transitional behavior further downstream, and second, because the blade curvature causes
significant damping or augmentation of turbulence transport in the boundary layers on the suction
and pressure sides, respectively. It is now generally accepted that the substantial variability in the
strength of the interaction between different strain types and the turbulent stresses can only be
resolved, in a fundamentally rigorous sense, through the use of second-moment closure, in which
separate transport equations are solved for all Reynolds-stress components. In particular, the very
different stress-generation terms contained in these equations give rise to that closure's ability to
resolve anisotropy and hence the influence of curvature, rotation and normal straining on the
stresses. However, this type of closure is complex, poses particular challenges in respect of its stable
integration into general computational schemes and is costly to apply in practice [Lien and
Leschziner]226. A simpler and more economical alternative, though one which rests on a weaker
fundamental foundation, is to use nonlinear stress/strain relations which can be made to return,
upon the introduction of physical constraints and careful calibration, some of the predictive
capabilities of 2nd moment closure.

226Lien, F.S., Leschziner, M.A. “Second-moment closure for three dimensional turbulent flow around and within
complex geometries”. Computers and Fluids 25, 237, 1996.
187

8.1 Case Study 1 - Turbulence Comparisons for a Low Pressure 1.5 Stage Test
Turbine
In a gas turbine engine, secondary flows have a unfavorable effect on efficiency. This phenomena
was investigated by [Dunn et al.]227. The current numerical study is aimed at determining which
turbulence model in a commercially available CFD code is best suited to predicting the secondary
flows. Experimental validation is used to determine the appropriateness of the model. It was found
that the Baldwin-Lomax, Spalart - Allmaras and κ-ε predicted the magnitude of the velocity well, but
did not capture the velocity magnitude profile well. The κ-ω and the SST κ-ω captured the profile
better, but did not predict the average value as well as the other models tested. As evident with
reference to the data, Baldwin-Lomax is still one of the best all-purpose models for turbomachinery.
The SST k-ω shows some promise as it predicted the velocity magnitude reasonably well, but the
Baldwin-Lomax predicted the radial and tangential velocities better. Due to the complex nature of
secondary flow it may still be sometime before computational hardware and the numerical models
are such that the complexities can be appropriately modelled. Until such time it is important that new
models be investigated and validated against experimental data. It is also recommended that the
computational expense be carefully weighed against the required accuracy.

8.2 Case Study 2 - Comparison of Various Turbulence Models in Rotating Machinery


Blade-to-Blade Passages
Authors : E.Y.K. Nga1 and S.T. TAN2
Affiliations : 1School of Mechanical and Production Engineering, Nanyang Technological University,
Nanyang Avenue, Singapore 639798. 2FerroTec Cooperation (S) Pte Ltd, Kallang Basin Industry Estate,
Kallang Ave., Singapore 639798
Citation : E. Y. K. Ng, S. T. Tan, "Comparison of Various Turbulence Models in Rotating Machinery Blade-
to-Blade Passages", International Journal of Rotating Machinery, vol. 6, Article
ID 385304, 8 pages, 2000. https://doi.org/10.1155/S1023621X0000035X
Source : International Journal of Rotating Machinery 2000, Vol. 6, No. 5, pp. 375-382
https://doi.org/10.1155/S1023621X0000035X
Adaptation : Slight Reduction to fit the format
Numerical calculations on four blade passages are done using Q3D Navier-Stokes solver with a simple
mixing length turbulence model and two more advanced transport-equation approaches. Mixing
length is simple and cheap but crude, while more sophisticated transport approaches are more
physical but more expensive. Predicted results using different turbulent models are discussed and
compared with the laminar flow and well documented experimental results. Studies show that the
model with more transport-equation predicts improved result as it includes the effects of upstream
history into the velocity scale.
8.2.1 Introduction
Most flow fields in fluid machinery passage are turbulent flow and it is one of the most complex
problem in the area of computational fluid dynamic such as stall and surge phenomena in compressor
system [14]. It is believed that the solution of time dependent three-dimensional full Navier-Stokes
equations could describe turbulent flows completely. However, the computers such as workstations
are not large and fast enough to solve the equations directly, for the required range of length and
time scales, even for simple flows. Hence, it is practical of using some of the turbulent modelling to
describe the turbulence motion instead of solving the full Navier-Stokes equation. Many publications
recommended various types of turbulent models such as those by Baldwin-Lomax [1], Cebeci-Smith

227Dwain Dunn, Glen Snedden, and T.W. von Backström, “Turbulence Model Comparisons for a Low Pressure 1.5
Stage Test Turbine”, ISABE-2009-1258.
188

[2], Birch [3], Chien-Kim [4], Launder-Spalding [5], Myong-Kasagi [6], etc.
Depending on the number of transport-equation used, the turbulent models can be classified into
zero-, one- and two-equation and higher-order models. Theoretically speaking, the more the number
of transport equations involved, the more
accurate the prediction is, as less assumptions
are used. The aim of this paper is to evaluate the
different types of turbulent models including
zero-, one- and two-equation models using a
Q3D Navier-Stokes [7] and full energy equation
unless otherwise stated, in one host code, with
experimental data for axial turbomachinery
application.
8.2.2 Turbulent Models
8.2.2.1 Baldwin-Lomax’s Zero-Equation
Model [1]
In zero-equation model, the concept of mixing
length is used. Dimensional analysis of variables
shows that the turbulent viscosity, μT, divided
by the density p has the same dimensions as a
length multiplied by a velocity. Hence
momentum arguments can be used to show that
μT is a function of the flow density, a length scale
in the flow and the local mean flow velocity. The
relationship is omitted here, while it is given in
[NG, E.Y.K. and TAN, S.T., “Comparison of Various
Turbulence Models in Rotating Machinery Blade-
to-Blade Passages”, 1999], and [Sadrehaghighi,
I., “Turbulence Modeling”, 2021] in full details.
8.2.2.2 Birch’s One-Equation Model [3]
In one-equation model, the turbulent viscosity
is related to the turbulent kinetic energy k.
Again, readers are encouraged to consult the
[NG and TAN, “Comparison of Various
Turbulence Models in Rotating Machinery Blade-
to-Blade Passages”, 1999] for additional
information.
8.2.2.3 Standard Two-Equation Model [5]
A standard k-ε model developed by Launder and
Spalding [5] is also coded. This is the most
commonly used model for CFD calculations. We
will avoid the formulation since it is available in
[NG and TAN, “Comparison of Various
Turbulence Models in Rotating Machinery Blade- Figure 8.2 (a) Grid generated for transonic
to-Blade Passages”, 1999],as well as, in compressor rotor (mesh: 86x 45); (b) Measured
[Sadrehaghighi, I., “Turbulence Modeling”, Mach contour at 45% span [9].
2021]. For I.Cs. and B.Cs. , users are encouraged
to consult the source as well.
189

8.2.3 Numerical Results


8.2.3.1 Transonic Compressor
Four experiment test cases are used to validate the predicted results. The first is a transonic
compressor rotor, experimental work has been done at the DFVLR [9] in Gottingen. The
computational H-mesh and measured Mach contour plot at 45% span are shown in Figure 8.2(a-
b). This mesh density has been determined to represent’ a good compromise between economy and
grid independence. It is sufficient to resolve down to either linear sublayer (for y+ ≤ 11.25) or log-
law layer (11.25 < y+ < 500) as the value of y+ for the grid next to the solid boundary is between 0.7
and 36.

Figure 8.3 Mach contour at rotor mid-span using (Left) Zero Equation (Middle) One Equation (Right)
Two Equation Turbulence Models

The Mach contours computed in Figure 8.3 show that, on the suction side, strong acceleration just
after the leading edge followed by a weak oblique shock. Different models produce slightly different
results. The shock predicted by zero equation model smears out into a wider number of grid.
However, both the one- and two-equation models predict a sharper shock which are closer to the
experimental result. All the models predict the shock at about 18-20% chord of the suction surface.
For the same axial chord, the one-
equation model produces the
highest Mach number, the zero
equation model gives the lowest
value while the two-equation model
predicts an in between value that is
closest to the experimental results.
In brief, the predicted result by each
model agrees qualitatively well
against the experimental data.
8.2.3.2 UTRS Turbine Blades
The next test cases are the UTRC
turbine blades. Experimental
results [10] have been widely
published. The blade surface
pressure-coefficient distributions
(Cp = (Pi - P)/O.5pi Ui2) based on the Figure 8.4 Blade pressure-coefficient prediction for UTRC
inlet condition for UTRC turbine turbine stator
stator at the nominal operating
190

point are compared in Figure 8.4. It shows similarity between the predictions using each turbulence
model. On the suction surface at same X/Cx chord position, the zero-equation model predicts a lower
value as compared to prediction by one- and two-equation models. However, on the pressure
surface, all the models predicted similar Cp values and are very close to the experimental result. As
the boundary layer at pressure surface is relatively thin and no significant separation occurs
therefore all the turbulent models are
able to predict more accurate results.
Included also is the Cp prediction by
laminar flow having similar shape to
turbulent assumption but with a much
lower Cp value at suction side.
Similarity, Figure 8.5 compares the Cp
distributions for turbine rotor. On both
the blade surfaces at same X/Cx position,
zero-equation model predicts a lower
value as compared to that by one- and
two equation models as well as the
experimental result. The latter models
are able to predict results which are very
close to experimental result on pressure
surface and second half of the suction Figure 8.5 Blade pressure-coefficient prediction for
surface. At the first 50% chord on the UTRC turbine rotor
suction surface, the experimental result
is lower than the prediction by one or two-equation models. Same as in case of stator the Cp
prediction by laminar flow has similar shape with that predicted by turbulent models but with a
much lower Cp value at suction side. In the measurement, the flow has passed through the stator,
which results in wake forming, flow distortion and non-uniformity at the rotor inlet in contrast to the
computation. Large error at entrance of rotor was also found in the calculation by Lee et al. [11].
8.2.3.3 C4 Compressor
The final case is on C4 compressor
blade where experimental data are
available from [12]. Figure 8.6
compares blade surface pressure
coefficient distributions at the
nominal operating point, showing
similarity between the predictions
using each turbulence model. On
suction surface, at same position, it
is shown that zero- and one
equation models predict similar
values which are slightly lower than
two-equation model as well as the
experiment values. However, on
pressure surface, zero and one
equation models predictions agree
well with measured data. Figure
8.6 also shows that the two- Figure 8.6 Blade pressure-coefficient prediction for C4
equation model agrees very well compressor blade
with the measurement at both
191

blade surfaces. The prediction of Cp for laminar flow is similar to that predicted by turbulent models
at most portion of the blade surfaces, except the region near trailing edge of the pressure side where
laminar model predicts a small
boundary separation bubble
and hence lower pressure
recovery as compared to
turbulent models.
Local velocity profiles along Y-
axis are plotted at 64% chords
and compare with experimental
result [12] as shown Figure
8.7. The Y/Ye term is the
nondimensionalized normal
distance from the blade surface
Y with the boundary layer
thickness Ye (defined as the grid
point with a speed less than
98% different compared to
adjacent grid). The speed U is
non-dimensionalized with Figure 8.7 Non-dimensionalized Y-direction distance vs. velocity
speed Ue which is the speed at curve at 64% chord of suction side of C4 compressor blade
Ye. At both chord-location, the
velocity profiles predicted by
two-equation model agree very well with the experiment. At above 20% of boundary layer thickness,
zero and one equation models predictions show good agreement with the experimental result.
However, below 20% of boundary layer thickness, zero equation model predicts lower value while
the one-equation model gives a higher value. In brief, all the models are able to predict the growth of
boundary layer and boundary separation. The laminar flow prediction is found to be quite similar to
the zero-equation model. More details could be found in [13].
8.2.4 Concluding Remarks
The computed results were compared with measurement to validate the code and assess the quality
of the numerical solution. The performance of the turbulence models to predict the flow through a
blade passage depends on the number of transport equations used and on the inlet flow conditions.
Another observation from the models used is their different separation behavior within the boundary
layers. It is shown that in most cases the two-equation model produces results which are closest to
the experimental results followed by one-equation model. As the less simplification is made the
closer to physics it will be. Finally, for accurate simulations of fluid machinery, extension to three-
dimension with transition and higher-order of turbulent model are needed.
8.2.5 References
[1] B. Baldwin and H. Lomax (1978). ’Thin layer approximation and algebraic model for separated
turbulent flows’, AIAA Paper, no. 78-257.
[2] T. Cebeci and A.M.O. Smith (1974). ’Analysis of turbulent boundary layers’, Applied Mathematics
and Mechanics, Vol. 15, Academic Press, New York.
[3] N.T. Birch (1987). ’Navier-Stokes predictions of transition loss and heat transfer in a turbine
cascade’, ASME 87-GT-22.
[4] Y.S. Chien and S.W. Kim (1987). ’Computation of turbulent flows using an extended k-c turbulence
model’, NASACR- 179204.
[5] B.E. Launder and D.B. Spalding (1974). ’The numerical computation of turbulent flow’, Comp.
Math. Appl. Mech. Eng., 3, 269-289.
192

[6] H.K. Myong and N. Kasagi (1990). ’A new approach to the improvement of the k-c turbulence
model for bounded shear flows’, JSME Int. J., Ser. B, 33, 63-72.
[7] E.Y.K. Ng and Y. Miao (1996). ’Viscous flow prediction of the single stage axial rotating
machineries’, CFD Journal, 403-420.
[8] H.K. Versteeg and W. Malalasekera (1995). An Introduction to Computational Fluid Dynamics,
Longman Scientific & Technical, Loughborough.
[9] L. Fottner (1990). ’Test cases for computation of internal flows in aero-engine components’, NASA
AGARD Advisory Report No: 275 (VI.4 Test Case E/CO-4 by Dunker, R.) pp. 245-285.
[10] R.P. Dring, H.D. Joslyn, L.W. Hardin and J.H. Wagner (1982). ’Turbine rotor and stator
interaction’, ASME Journal of Engineering for Power, 104, 729-742.
[11] Y.T. Lee, T.W. Bein, J. Feng and C.L. Merkle (1993). ’Unsteady rotor dynamics in cascade’, ASME
Journal of Turbomachinery, 115, 85-93.
[12] E.Y.K. Ng (1992). ’A high resolution coupled parabolic/elliptic Navier-Stokes solver for
turbomachinery flows’, Ph.D. Thesis, University of Cambridge, UK.
[13] S.T. Tan (1998). Numerical investigation of viscous flow in rotating cavity with different
turbulence models, M. Eng. Thesis, Nanyang Technological University, Singapore.
[14] E.Y.K. Ng, S.T. Tan and H.N. Lim (1997). ’Simulation of instability flow in compressor system’,
Proceedings of 8th Aerospace Technology Seminar, Singapore, pp. PR 4/1-4/7.
193

9 Rotor-Stator Interaction Treatment (RST)


9.1 Physical Perspectives
Turbomachinery flows are naturally unsteady mainly due to the relative motion of rotors and stators
and the natural flow instabilities present in tip gaps and secondary flows228. Full scale, time
dependent calculations for unsteady turbomachinery flows are still too expensive to be suitable for
daily design purposes. One of the reasons for this large cost is the fact that in practical
turbomachinery of these reduced order models requires that the engineer/designer be aware of a
method's capabilities as well as its limitations. The key trade off in the computation of unsteady
turbomachinery flows is between the accuracy of the method and the cost or computational
efficiency with which a solution can be obtained. Highly accurate and well resolved models tend to
be limited by the available computing power, while most reduced-order models usually neglect a
significant amount of the physics and are therefore not credible for the evaluation of the performance
and heat transfer characteristics of a turbomachine. A balance between these extremes is clearly
desirable. In order to include the unsteady effects while keeping the computational requirements
reasonable, two types of approximations can be distinguished. The first approach involves rescaling
the geometry (typically by altering the blade counts and their chords to maintain solidity) such that
periodicity assumptions hold in an azimuthal portion of the domain that is much smaller than the full
annulus. A second alternative involves the use of the original geometry but compromises the fidelity
of the time integration method229. Figure 9.1 shows a schematics of 3D point of view in either case.

Figure 9.1 Schematics of 3D Concept at IGV/Rotor/Stator Interface

228 Xiaodong Wang, “CFD Simulation of Complex Flows in Turbomachinery and Robust Optimization of Blade
Design”, Submitted to the Department of Mechanical Engineering Doctor of Philosophy at the Vrije Universiteit
Brussel July 2010.
229 Arathi K. Gopinath, Edwin van der Weidey, Juan J. Alonsoz, Antony Jamesonx, Stanford University, Stanford,

CA 94305-4035, Kivanc Ekici {and Kenneth C. Hallk, Duke University, Durham, NC 27708-0300, “Three-
Dimensional Unsteady Multi-stage Turbomachinery Simulations using the Harmonic Balance Technique”
194

All of these approximations can be considered to be


different variations of reduced-order models. In Steady Un-Steady
order to use the same solver, the flows in stator and Mixing Plane Sliding Mesh
rotor should be calculated in the stationary frame of Frozen Rotor Harmonic Balance
reference and the rotating frame of reference, Time Transformation
respectively. However, a critical problem is how to
transfer the information downstream and upstream Table 9.1 Rotor/Stator Interaction
at the interface of stator and rotor. The quality of the Schemes
flow predictions for multistage turbomachinery
strongly depends on the treatment of rotor/stator interaction. Figure 9.2 illustrates the interface
between them. Two general approaches as steady and unsteady interactions are available as detailed
in Table 9.1.

Figure 9.2 Interface Between Rotor/Stator

9.2 Multi-Passage vs. Multi-Stages


Before going any further, it is worth mentioning two terminology which is been used often in
literature. They are Multi-
Stage and Multi-Passage. The
difference been best explain
through the Figure
9.3. It could thought of a matrix
notation. While passages are
the column of matrix and
usually treated the same in
terms of gridding and analysis,
the rows are stages and treated
another way as they have
usually different geometry and
conditions.
Adjacent blade rows contain
unequal numbers of blades and Figure 9.3 Difference between Passage and Stages
shape, therefore, in principle, a
195

proper simulation requires solution of all blades in each row. However, some vendors such as
ANSYS© has developed a suite of tools that enables more efficient solution for a number of analysis
types. The key attribute of these tools is that the full wheel solution can be obtained by solving only
one or at most a few blades per row230.

9.3 Case for Mixing Plane Model


The quasi-steady methods based on mixing models which have been widely applied to flow
computations of multi-stages turbomachinery, such as the work done by [YaLu et al.]231 , [Pengcheng
& Fangfei]232 , and [Wang]233. In the interim, the unsteady numerical simulation has also been used
due to its ability in obtaining time-dependent flow solutions. Through-flow method, mixing model,
passage-averaging and unsteady computation are the typical approaches for numerical simulation of
multi-stage turbomachinery flow. The through-flow method, proposed by [Wu]234 , decomposes the
three-dimensional turbomachinery flow into a pair of two-dimensional flows, by which the flow
solutions can be iteratively obtained. The mixing model method, first proposed by [Denton]235,
employs an interface, called mixing plane, between adjacent blade rows, where the circumferentially
averaged flow solutions on each side are exchanged. This method transfers the inherent unsteady
flow of multi-stages into a quasi-steady one, resulting in a balance between the significantly reduced
computational cost and the fidelity of flow solutions. However, in the situations with plenty of blade
rows or small axial gap between adjacent blade rows, the flow solutions by mixing models deviate
from the experiments, or even no converged results can be obtained.
The passage-averaging method, developed by [Adamczyk]236 , transfers the unsteady flow into the
time-averaged one in a single blade passage by introducing three averaging operators in the Navier-
Stokes equations. However, it is rarely applied due to the complexity and difficulty to close the
correlated terms. By solving the unsteady Euler, unsteady Reynolds-averaged Navier-Stokes
(URANS) equations, more flow details can be obtained. Furthermore, the turbulence flow in
turbomachinery is rather complex and demonstrates a strong non-equilibrium turbulent transport
nature, it is still a challenge for predicting the flow correctly. Hybrid LES/RANS could give more
reasonable results and can be used to investigate the flow mechanism, but with enormous
computational cost. At present, the mixing model method is the most popular one in the flow
computations of multi-stages. The crucial issue of mixing model method is the appropriate model
used to simulate the flow mixing process on the interface between adjacent blade rows. In order to
match the flow mixing process as far as possible, a practical mixing model should be able to :

230 Turbomachinery Simulation, ANSYS blog.


231 ZHU YaLu, LUO JiaQi & LIU Feng, “Flow computations of multi-stages by URANS and flux balanced mixing
models”, Science China, Technological Sciences, July 2018 Vol.61 No.7: 1081–1091.
232 Du Pengcheng and Ning Fangfei, “Validation of a novel mixing-plane method for multistage turbomachinery

steady flow analysis”, Chinese Journal of Aeronautics, (2016), 29(6): 1563–1574.


Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row
Turbomachines”, Siemens Industrial Turbomachinery Ltd., Waterside South, Lincoln LN5 7FD, UK.
232 Wu C H. “A general theory of three-dimensional flow in subsonic and supersonic turbomachines of axial, radial,

and mixed-flow types”, NASA TN 2604, 1952.


232 Denton J. D. “The calculation of three dimensional viscous flow through multistage turbomachines”, 1990.
232 See 148.
233 Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row

Turbomachines”, Siemens Industrial Turbomachinery Ltd., Waterside South, Lincoln LN5 7FD, UK.
234 Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row

Turbomachines”, Siemens Industrial Turbomachinery Ltd., Waterside South, Lincoln LN5 7FD, UK.
235 Denton J. D. “The calculation of three dimensional viscous flow through multistage turbomachines”,1990.
236 Adamczyk J J, Mulac R A, Celestina M L. “A model for closing the inviscid form of the average-passage

equation system”. ASME, 1986. 86-GT-227.


196

➢ Keep strong conservation of flow solutions across the interface, such as mass flow rate,
momentum, total enthalpy, etc.;
➢ Obtain the aerodynamic parameters with little deviation from the experiment;
➢ Be robust by employing non-reflective boundary conditions on the interface and special
interface treatments for reversed flow.

A popular class of simple mixing models simulate the flow mixing process by simply circumferentially
averaging the flow variables on the interface. Since there are only five independent flow variables in
three-dimensional compressible flow, different selections of the independent flow variables result in
different simple mixing models. In the past decade, some novel mixing models have also been
proposed. All the results demonstrate that different mixing models have various effects on the
computation robustness, flow solution conservation and thus the flow details. However, no
comparative investigation of the aforementioned novel mixing models has been carried out in the
open literatures. With the development of computer capacity, more emphases are put on URANS. The
unsteady computation methods for multi-stages include the phase-lagged method, blade scaling
technique, time-inclined method and frequency domain methods, such as nonlinear harmonic
method and harmonic balance method. Due to the eases of implementation and extension to multi-
stages, the blade scaling technique has been widely applied in the unsteady turbomachinery flow
computations. The exchange of two-dimensional flow fields on the interface between adjacent blade
rows is the most crucial issue for unsteady flow computation of multi-stages because of the non-
matched grid points between the two sides of the interface and the relative motion between rotor
and stator. In such cases, it is necessary to develop an interpolation method strongly maintaining the
conservation and continuity of flow variables across the interface.

9.4 Steady Treatment of Interface (Mixing Plane)


The simplest treatment of R/S interface is the stage or Mixing Plane method proposed by [Denton]237.
This method assumes the exiting flows of stator become uniform flows before entering the inlet of
domain of rotor. A block computational domain of Rotor, Guided Vanes, Mixing Planes and applied
boundary is shown in Figure 9.4. A pitch wise averaging of the flow solution is needed at R/S
interface before transferring the information of
both sides. The essential idea behind the
mixing plane concept is that each fluid zone is
solved as a steady-state problem238. At some
prescribed iteration interval, the flow data at the
mixing plane interface are averaged in the
circumferential direction on both the stator
outlet and the rotor inlet boundaries. The
averaging process could be choice of three types
of averaging methods: Area-weighted averaging,
Mass averaging, and Mixed-out averaging. By
performing circumferential averages at
specified radial or axial stations, "profiles'' of
boundary condition flow variables can be
defined. These profiles, which will be functions Figure 9.4 Block Computational Domain for a
of either the axial or the radial coordinate, Rotor with guiding Vanes

237 J. D. Denton, “The calculation of three-dimensional viscous flow through multistage Turbomachinery”,
Journal of Turbomachinery, 114(1):18–26, 1992.
238 Release 12.0 © ANSYS, Inc. 2009-01-22.
197

depending on the orientation of the


mixing plane, are then used to update
boundary conditions along the two
zones of the mixing plane interface. In
the examples shown in Figure 9.5
profiles of averaged total pressure (P0),
direction cosines of the local flow
angles in the radial, tangential, and axial
directions (αr, αt, αz), total temperature
(T0), turbulence kinetic energy (k), and
turbulence dissipation rate (ε) are
computed at the rotor exit and used to
update boundary conditions at the
stator inlet. Likewise, a profile of static
pressure (Ps), direction cosines of the
local flow angles in the radial, Figure 9.5 Axial Rotor/Stator Interaction (Schematics
tangential, and axial directions (αr, αt, Illustrating the Mixing Plane concepts)
αz), are computed at the stator inlet and
used as a boundary condition on the rotor
exit. Note that the meshes on both sides of
the interface should cover the same range
in span wise, the averaging is performed
along the same azimuthal mesh lines.
However, a full non-matching mixing
plane239 can be used to overcome this
limitation. Better, the isolated simulation
on single stator or rotor, the interaction of
potential flows in considered in this
method. However, the impact of secondary
flows and separation flow are erased. This
physical approximation tends to become
more acceptable as rotational speed is
increased. The mixing plane method is by
far the most often used R/S modeling in
industry design and optimization.
Figure 9.6 A Compressor Pressure Distribution on a
Unfortunately it doesn’t capture the whole
Surface using a Mixing Plane
physics. This is usually evident by visual
inspection of in interface (mixing) plane as
an imaginary line between the cascades. Figure 9.6 displays a compressor Pressure Distribution on
a surface at constant radius half way between the hub and the casing using a Mixing Plane
computation.
9.4.1 Losses Across the Interface of Mixing Plane
In a CFD-based steady flow field analysis for a multiple-blade-row turbomachine, one blade passage
is usually used for one blade row, and there is an artificial interface, also called mixing plane,
between adjacent blade rows240. The one-to-one correspondence of corresponding points on such an
interface and in two adjacent blade-row domains is lost completely in such an analysis. This raises

NUMECA International, Brussels, “Fine/Turbo User Manual V8”, October 2007.


239
240Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row
Turbomachines”, Siemens Industrial Turbomachinery Ltd., Waterside South, Lincoln LN5 7FD, UK.
198

the issue of how to transfer solution information across such an artificial interface in a steady flow
field analysis.
In reality, upstream wake mixes out gradually when it is transported downstream. The mixing loss is
also expected to rise gradually. With a mixing-plane treatment, the strong circumferential non-
uniformity of an upstream wake mixes out significantly across an interface, leading to a nearly
circumferential uniform flow field on the downstream side. Consequently, across an interface from
its upstream side to its downstream side there is an abrupt rise of loss. Independent research shows
that the artificial mixing loss across the interface of a turbine stage by a steady mixing-plane analysis
is significant, thus leading to higher overall loss for the turbine stage in comparison with the loss
from an unsteady sliding-plane analysis. However, some research demonstrates that the artificial
mixing loss can be trivial, leading to lower overall loss than that by an unsteady sliding-plane analysis.
Nevertheless, investigations by both [Fritsch and Giles]241 and [Pullan]242 indicate that the loss by a
steady mixing-plane analysis grows at a slower rate through the downstream blade row in
comparison with that by an unsteady sliding-plane analysis. Modern multiple-blade-row
turbomachines usually have a small inter row gap. This situation makes it not only desirable but also
necessary for a mixing-plane method to be non-reflective. Otherwise artificial reflections from a
mixing plane very close to a blade leading edge or a blade trailing edge can deter convergence and
spoil solution. Apart from conservation and non-reflectiveness, as pointed out in243, an ideal mixing-
plane method should also be robust so that it can handle reverse flow, which can exist either in a
solution process or in a converged flow field.
9.4.2 Principles of Flux Conservation
An artificial interface between two adjacent blade rows is normally a revolution surface, which is a
single curve connecting the hub contour and the casing contour in the meridional plane as shown in
Figure 9.7244. The conservation law states that the fluxes of mass, momentum, energy, and other
scalar quantities through an arbitrary segment of the interface from the domain on one side of the

Figure 9.7 Schematic of an Artificial Interface Between a Rotor and a Stator (left) and the Virtual
Control Volume Formed by Displacing Two Adjacent Domains (right)

241 Fritsch, G., and Giles, M. B., “An Asymptotic Analysis of Mixing Loss,” ASME J. Turbomachines, 1995.
242 Pullan, G., 2006, “Secondary Flows and Loss Caused by Blade Row Interaction in a Turbine Stage,” ASME J.
Turbomachine., 128, pp. 484–491, July 2006.
243 Holmes, D. G., “Mixing Planes Revisited: A Steady Mixing Plane Approach Designed to Combine High Levels of

Conservation and Robustness,” ASME Paper No. GT2008-51296, 2008.


244 Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row

Turbomachines”, Siemens Industrial Turbomachinery Ltd.,


Waterside South, Lincoln LN5 7FD, UK.
199

segment should be equal to those into the domain on the other side of the segment. This can be easily
understood through forming a virtual control volume, the hatched part as shown in Figure 9.7, by
displacing the two adjacent domains in the stream wise direction. There is no flux through the upper
and lower sides of the control volume; therefore, the flux entering the control volume from one
domain must equal the corresponding flux exiting the control volume into the other domain in a
steady flow analysis. For an arbitrary segment with the length of ds as shown in Figure 9.7, the
fluxes of mass, momentum, energy, and scalar quantities through the segmental revolution surface
are defined as follows:

Mass Flux: F1 = ∫ ρvn dsrdθ


0

Axial Momentum Flux: F2 = ∫ (ρvn vx + p. nx )dsrdθ


0

Tangential Momentum Flux: F3 = ∫ (ρvn vx + p. nx )dsrdθ


0

Radial Momentum Flux: F4 = ∫ (ρvn vr + p. nr )dsrdθ


0

Total Energy Flux: F5 = ∫ ρvn H dsrdθ


0

Arbitary Scalar Quatitity F6 = ∫ ρvn ∅ dsrdθ


0
Eq. 9.1
The area of the segmental surface is given by

S = ∫ dsrdθ
0
Eq. 9.2
Dividing Eq. 9.1 by this area gives the circumferential area averaged fluxes. It is obvious that, for a
fully converged steady solution of a multiple-blade-row turbomachine, circumferential area
averaged fluxes of mass, momentum, energy, and other scalar quantities across an arbitrary segment
of an interface, as defined in Eq. 9.1/Eq. 9.2, should be conserved. It should be noted the
circumferential area averaged fluxes as defined in Eq. 9.1/Eq. 9.2 can be calculated over a pitch
angle that is less than 2π. Before a solution converges, the fluxes calculated using flow variables from
domains on its two sides of an interface are usually not conserved or equal. The task of a mixing-
plane method is to make use of the fluxes to drive the differences to zero. Two existing methods will
be explained with the proposition of a new method later. To get the flow primitive variable
200

differences, 3 methods is been proposed by [Wang]245. The differences of the three methods lie in
how to calculate the incoming flow disturbances according to circumferential area-averaged flux
terms on two sides of an interface. For further information, please consult 246.
9.4.3 Case Study 1 - Comparison of Flux
Balanced Mixing Models on Q-1.5 Steady Mixing Model Nomenclature
Stage Rotor 67 Momentum-Averaged MA
To study the performance of flux balanced Entropy-Averaged EA
mixing models on simulating the complex Flux Balanced MA
flow, such as shock wave and reversed (FBMA)
flow on the interface, the quasi-1.5-stage Flux Balanced
Flux
Rotor 67 with small geometric Balanced EA (FBEA)
modification is then investigated . An Experiment:
artificial step of about 6% span is imposed TA
Time-Averaged
on the casing contour ahead of the second Experiment:
interface to produce reversed flow on the MP
Total Pressure Ratio
interface. The designed and modified
casing contours are shown in Figure 9.8. Table 9.2 Nomenclature for different Mixing Models
The grid with the same topology and cell Used in Study
number used for flow solver validation is
generated. Table 9.2 lists the nomenclature for Mixing models for steady computation. For a listing
of relative deviations of flow variables across the 1st and 2nd interfaces of the modified quasi-1.5
stage Rotor 67 vs. the circumferential velocity, see [YaLu et al.]247. On both interfaces, the unsteady
computation has an absolute
advantage on the flow
conservation over all the
mixing models. Compared
with the simple mixing models,
the relative deviations of most
of flow variables are decreased
by the flux balanced models.
By the flux balanced models,
the total temperature ratio
across all the interfaces is
almost strictly maintained,
demonstrating the superior
performance of flux balanced
mixing models on flow
conservation.
Compared with other flow
variables, the relative Figure 9.8 Sketch of Casing Treatment of Rotor 67 (Courtesy of 157)
deviation of tangential velocity
across the first interface is extremely large, especially for MA and EA models. This is because the
tangential velocity on the interface is quite small. The entropy by MA and EA models decreases across

245 Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row
Turbomachines”, Siemens Industrial Turbomachinery Ltd., Waterside South, Lincoln LN5 7FD, UK.
246 See Previous.
247 ZHU YaLu, LUO JiaQi & LIU Feng, “Flow computations of multi-stages by URANS and flux balanced mixing

models”, Science China, Technological Sciences, July 2018 Vol.61 No.7: 1081–1091.
201

the first interface,


violating the physical rule of
entropy production. It is
supposed to be induced by the
large deviation of tangential
velocity. The total pressure ratio
by FBEA model slightly increases
across the first interface, also
violating the physical rule. It is
supposed to be induced by the
slightly increased entropy
production across the interface.
Figure 9.9 shows the span-wise
distributions of total pressure
ratio and total temperature ratio
on the first interface. The span-
wise distributions of total
pressure ratio by the mixing
models perform rule-less Figure 9.9 Span-Wise Distributions of Aerodynamic Parameters
variations, whereas the on the 1st Interface of Modified Rotor 67- (a) Total Pressure Ratio;
deviations from time-averaged (b) Total Temperature Ratio (YaLu et al.)
ones are slight. The span-wise
distributions of total temperature ratio by the simple mixing models are far away from the time-
averaged distribution, whereas the ones obtained from flux balanced mixing models are almost the
duplicates of time-averaged distribution. The discrepancies of flow variables among the present
methods on the middle and upper spans are associated with the shock wave. The shock wave
originated from the leading edge of rotor blade injects onto the first interface, which can be
illustrated by the contour of relative Mach number at 50% span in Figure 9.10. The position of
shock wave on the interface can also be clearly displayed by the contour of static temperature in
Figure 9.11. The shock wave injects onto the interface from 20% to 75% spans, where the span-

Figure 9.10 Contour of Relative Mach Number and Iso-Surface of Axial Velocity of Modified Rotor 67
202

wise distributions of flow variables by different computation methods are not consistent with each
other as shown in Figure 9.9. Although the positions of shock wave on the interface are almost the
same for all the methods, the detailed shock wave patterns are slightly different as indicated by the
zones with low static temperature in Figure 9.11. The shock wave patterns by MA and EA models
are similar, and those by FBMA and FBEA models are also close, which are consistent with the span-
wise distributions of total pressure ratio and total temperature ratio. However, none of shock wave
patterns by the mixing models matches well with that of unsteady computation. The patterns of
rarefaction waves after the shock waves perform the similar variations.

Figure 9.11 Temperature Contours on the 1st Interface of Modified Rotor 67 - (a) FBEA ; (b) FBMA ; (c)
EA ; (d) MA ; (e) TA (Courtesy of YaLu et al.)

9.4.4 Case Study 2 - Modeling of Secondary Flows in Single Blade Rows using Mixing Plane
Approach
A computational modeling of secondary flows in single blade rows and a performance assessment in
3D turbine stages computations using wall functions were made by [Xisto et al.] 248 . The analysis of
the flow in turbine blades has been extended from 2D to 3D, and from pure Euler equations to Navier-
Stokes modeling, including turbulent flow. This later has only been possible due to fast development
of computer power of modern desktop computers. Most of these analyses were carried out for
isolated blade rows. However, this approach is not accurate in many circumstances, due to a strong
coupling and interaction between the several blade rows. To fully account for the rotor-stator
interaction, a 3D unsteady Navier-Stokes analysis is required, but such an analysis is too CPU-
intensive and expensive in terms of computing power, so we will restrict our approach to the mixing
plane model . The mixing plane approach is applied at the blade row interface between the stator-
rotor. It can also be applied to several stages in series. In this approach one assumes that the flow is
totally mixed out and is axis-symmetric between the blade rows. Actually, it can only include the
effects of radial variation in an approximate way and cannot account for any circumferential
variations, such as those created by wakes, leakage or secondary flows. Although this, it is important
to clarify that the pitch wise averaging does not affect the span wise variation in flow. Actually, the
span wise variation of pressure, velocity, flow angle, etc., at all stations between hub and tip is
obtained from the full 3D Navier-Stokes computation.
The computation of the flow, for a single blade row, can nowadays routinely be made using a low-Re

248Carlos M. C. Xisto , José C. Páscoa e Emil Göttlich, “Computational modeling of secondary flows in single blade
rows and performance assessment in 3D turbine stages computations using wall functions”, November 2009.
203

turbulence model and resolving the boundary layer, even with desktop computers. The computation
of a whole stage is more computing demanding and, at least with our current capabilities, can only
be accomplished with the use of wall functions. This introduces the reason for the current work,
which is to analyze the performance assessment limitations when using wall functions, for turbine
stage computations, instead of resolving the boundary layer. A major problem that arises in the
design and performance analysis of axial turbines is the understanding, analysis, forecasting and
control of secondary flows . A pioneer work in the understanding of this phenomenon could be found
in . Here Langton presents the evolution of a tree-dimensional flow in a turbine cascade. Were at the
end wall of the cascade the inlet boundary layer separates at the saddle-point and forms the
horseshoe vortex. The pressure leg of this vortex will become the passage vortex and the suction leg
will become the counter-vortex and as an opposite sense of rotation to the passage vortex, see
Figure 6.11.
9.4.4.1 Transonic Turbine Stage Meshing and Flow Details
A high level of detail for the geometry was considered, including the fillets at hub and tip sections and
the rotor gap. The stator comprises 24 blades and the rotor has 36 blades, which represents a ratio
of 2:3 between the stator and rotor blades. In our case, using the mixing plane model we can perform
the computation using only one blade for the stator and rotor rows. The first phase of the
computations performed for the stage was made using isolated blade rows for the stator and rotor.
By solving each flow in an isolated blade row we were able to detect any flow convergence problems,
typically created by poor mesh quality. After this fine tuning of the mesh we proceed into the full
stage computation. For this test case we solved the Navier-Stokes equations using the Spalart-
Allmaras turbulence model. In this computation an implicit discretization using double precision
was retained. The computations started using a first-order discretization in space and later on were
toggle to second order accuracy. The mesh comprises 15 H blocks, with 8 blocks in the stator and 7
in the rotor. The overall mesh comprises 224136 nodes, these were 131136 for the stator and 93000
for the rotor. The stator blade
comprises 49 points in the
inter blade region and 93
points in the axial flow
direction, with 65 points used
to define the blade surface.
For the radial direction we
have distributed 30 points.
For the rotor blade 29 points
were applied in the inter-
blade zone and 93 on the axial
direction, with 65 points used
to define the blade geometry.
In the radial direction 30
points were used, the mesh
can be seen in Figure 9.12.
The flow field at inlet of the
stator is completely subsonic,
transonic flow is restricted to
minor zones around the
stator trailing edge. Thus, at Figure 9.12 Mesh for Transonic Turbine Stage - Upper Image
stage inlet we have imposed Depicted the Mesh at the Hub Surface while the Lower Image
stagnation pressure and Represented Mesh used for the Blade Span
temperature with the
204

corresponding flow angles and at stage outlet static pressure is imposed. At the mixing plane
interface also characteristic boundary conditions are imposed, namely stagnation pressure,
temperature, flow angles at rotor inlet and static pressure at stator outlet. In order to apply the
Spalart-Allmaras turbulence models we have considered a turbulence intensity of 10% and a length
scale of 1% pitch at stator mid span. These turbulence quantities are usually applied in modeling
turbomachinery flows. The initial computations were performed with an explicit approach and using
pure characteristic boundary conditions to extrapolate the variables at the boundaries.
Unfortunately convergence was
not attained, the residues got
stuck at a minor value.
Convergence was attained only
when using the implicit
formulation and applying
nonreflecting boundary
conditions. Due to computing
power restrictions only results
for the Spalart-Allmaras
turbulence model using wall
functions were obtained. Figure
9.13 presents the numerical
results of velocity obtained for a
section at stator mid span.
Although, numerical result
shows a good agreement with the Figure 9.13 Results of the Velocity Contours for a Radial Section
experimental data, however, at Stator Mid Span using the Mixing Plane Approach
more experimental data is
necessary for a precise validation of the model. Only with a span wise distribution of experimental
and computed variables we can assess, in full, the capability of the mixing plane model in the
prediction of this flow field. For further details, readers are encourage to consult [Xisto et al.]249 .
9.4.5 Case Study 3 - Improvement Methods for Mixing Plane Models
For modern turbomachines, the trend of design is to reach higher aerodynamic loading but with still
further compact size. In such a case, the traditional mixing-plane method has to be revised to give a
more physically meaningful prediction. [Pengcheng & Fangfei]250, presented a novel mixing-plane
method, and three representative test cases including a transonic compressor, a highly-loaded
centrifugal compressor and a high pressure axial turbine were performed for validation purpose.
This novel mixing-plane method can satisfy the flux conservation perfectly. Reverse flow across the
mixing-plane interface can be resolved naturally, thus making this method numerically robust.
Artificial reflection at the mixing-plane interface is almost eliminated, and then its detrimental impact
on the flow field is minimized. Generally, this mixing-plane method is suitable to simulate steady
flows in highly-loaded multistage turbomachines.
From the authors’ point of view, the mixing-plane method should not just be a pure numerical
procedure to transfer the circumferentially averaged flow variables across the interface. A physical
correspondence for this pitch wise mixing can be found, i.e., we can just make the gap between the
two adjacent blade rows long enough so as to mix out all the non-uniformities (as shown from Figure

249 Carlos M. C. Xisto , José C. Páscoa e Emil Göttlich, “Computational modeling of secondary flows in single blade
rows and performance assessment in 3D turbine stages computations using wall functions”, Conference Paper·
November 2009.
250 Du Pengcheng a, Ning Fangfei, “Validation of a novel mixing-plane method for multistage turbomachinery

steady flow analysis”, Chinese Journal of Aeronautics, (2016).


205

9.14-(a) and (b)), while the span wise mixing is assumed to be suspended in the ‘‘extended mixing
region”. Therefore, for a fully converged flow field in the case as shown in Figure 9.14-(b), we can
find such an intermediate position where the flows are pitch wise uniform. At this position, if we cut
out an infinitely thin slice as denoted by two lines ‘‘ml” and ‘‘mr” in Figure 9.14-(b), the following
governing equations expressed in cylindrical coordinate system hold:

̅
∂𝐐
= 𝐅mr − 𝐅ml , ̅ = [ρ, ρv, ρv, ρv, ρe]T
𝐐
∂t
𝐅 = [ρU, ρUvx + nx p , ρUvθ + nθ p , ρUvr + nr p , ρUH]T
Eq. 9.3
with t being the pseudo time, ρ the density, (vx; vɵ; vr) the absolute velocity components expressed
in cylindrical coordinate (x , ɵ , r), e the total energy, U = nxvx + nɵvɵ + nrvr the advective velocity
normal to the blade row interface, p the static pressure and H the total enthalpy. The unit vector n =
(nx , nɵ , nr)T denotes the normal direction of blade row interface, and n is actually equal to zero
because the interface is a revolution surface.

Figure 9.14 Schematic View of Pitch-Wise Mixing Model

9.4.5.1 Validation Test Case


Solving the integral form of the governing equations which are discretized in space using a cell-
centered finite-volume method. The advective fluxes are evaluated using the low-diffusion flux-
splitting scheme coupled with Monotone Upstream-Centered Schemes for Conservation Laws
(MUSCL) interpolation to obtain high-order spatial accuracy. The diffusive fluxes are solved using
traditional central differencing. The one-equation Spalart–Allmaras turbulence model is used for
turbulent flows, which is discretized and solved in a coupled manner with the mean flow equations.
Message Passing Interface (MPI) is used to parallelize the code. The discretized system is solved
using the so-called matrix-free Gauss–Seidel algorithm.
9.4.5.2 1.5 Stage Transonic Axial Compressor
A 1.5 stage compressor extracted from a multistage high pressure compressor is considered first. In
order to perform fast unsteady calculation as reference, the blade count number ratio is scaled to be
1:1:2. In the unsteady simulation, the rotor shock travels upstream across the first mixing-plane
(MP1) due to the high loading of the rotor and relatively small gap from the upstream inlet guide
vane (IGV) (Figure 9.15-(A)). Meanwhile, the thick rotor wakes propagate through the second
mixing plane (MP2) and interact with the downstream stator (Figure 9.15-(B)). Thus, this test case
206

is very suitable for demonstrating the effectiveness of the proposed mixing-plane model in dealing
with the typical circumferential non uniform flow field featured by strong shock and thick wakes.
Further details can be obtained from .

(A) Static
(B) Entropy
Pressure

Figure 9.15 Instantaneous Distributions at 90% Span.

9.4.6 Frozen Rotor


If the exchange of information at the interface is by interpolation directly without averaging, one has
the Frozen Rotor method. The difference is, Mixing Plane mixes the flow and apply the average
qualities on the interface for upstream and downstream components; while frozen rotor will
pass the true flow to down steam and vice versa. So if you are interested in the wake effect on the
downstream component performance then you should use frozen rotor method. Its disadvantage is
that, if gives you the solution at the single relative position. So if you want to get the wake effect on
the downstream component for all relative positions (as happens in reality) then you should go for
the true transient method. As the name indicates, the relative position of rotor and stator is fixed.
Hence, the result of the frozen rotor method is equivalent to a certain point of the unsteady
simulation which means the flow solutions will dependent on the relative position between rotor and
stator. Since the information exchange on R/S interface is through interpolation, the mesh on both
sides of the R/S interface should cover the same pitch range. That means the periodic of the rotor
domain and stator domain should be kept the same,

K S PS = K R PR Eq. 9.4

Where, KS and KR are relative prime which stand for the number of passages in the stator domain and
rotor domain, respectively. Ps and PR denote the pitch of stator and rotor separately. An
approximation of the blade number can be made if Ks and Kr are large in order to reduce the
computational cost, which is called Domain Scaling. For instance a turbine with 29 blades of stator
and 31 blades of rotor can be approximated by a turbine with 30 blades of both stator and rotor, then
only one passage is needed to mesh for both stator and rotor. However, the simulation results are
only the approximated result to the real model. The Frozen Rotor method is used firstly by [Brost et
al.]251 in simulations of an axial turbine where the simulated results have a good accordance with the
transient results of the measurement. While, the flow field in a passage usually changes a lot during

V. Brost, A. Ruprecht, and M. Maih, “Rotor-Stator interactions in an axial turbine, a comparison of transient
251

and steady state frozen rotor simulations”, Conference on Case Studies in Hydraulic Systems-CSHS03, 2003.
207

the period. Therefore, this method


is only used in some specific
simulations. The information
exchange processes of mixing
plane and frozen rotor methods
depend on the boundary type of
the R/S interface. The detail
settings for different boundary
types and the corresponding
exchange strategies can be found
in252. (See Figure 9.16).

9.5 Un-Steady Treatment of


Interface
9.5.1 Sliding Mesh (MRF)
Full unsteady simulations that
integrate the governing equations
in time can be performed to model
the nonlinear unsteady Figure 9.16 Total Pressure Calculated by the Frozen Rotor
disturbances by marching time
accurately from one physical time
instant to the next. The flow fields within multiple blade rows are solved simultaneously and the
meshes within adjacent rows are moved relative to one another with each time step. However, the
computational expense of this approach can be significant. This is because sub-iterations are
required at each time instant, the time step size is necessarily small to preserve time accuracy, and
many time steps are required to reach a time periodic solution. Additionally, multiple passages must
be meshed to achieve spatial periodicity, unless so-called phase-lagged boundary conditions are used
to reduce the size of
the computational domain to a single blade passage in each blade row253. For unsteady simulation, a
natural idea is to simulate several different transient positions of rotor related to stator which leading
to the traditional unsteady treatment of R/S interface is the Sliding Mesh method proposed by Rai254.
For unsteady simulation, a natural idea is to simulate several different transient positions of rotor
related to stator which leading to the traditional unsteady treatment of R/S interface is the Sliding
Mesh method proposed by [Rai]255. In this method, the computational domain is divided into two
parts: rotor domain and stator domain. The mesh for rotor domain rotates with rotor. The R/S
interface becomes a sliding face and the exchanges of solution information are through the
interpolation to the dummy cells on both side without any averaging. At each time step, the rotor is
set at its correct position and equations are solved for that particular time step for the whole
computation domain. The final solution is therefore a succession of instantaneous solutions for each
increment of the rotor position. More precisely to set up sliding mesh simulation256,

252 NUMECA International, Brussels, “Fine/Turbo User Manual V8 (including Euranus)”, October 2007.
253 J. M. Weiss, K. C. Hall, “simulation of unsteady turbomachinery flows using an implicitly coupled nonlinear
harmonic balance method”, Proceedings of ASME Turbo Expo 2011, GT2011.
254 M. Rai, “Application of domain decomposition methods to turbomachinery flows”, ASME Advances and

Applications in Computational Fluid Dynamics, volume 66, 1988.


255 M. Rai, “Application of domain decomposition methods to turbomachinery flows”, ASME Advances and

Applications in Computational Fluid Dynamics, volume 66, 1988.


256 Reza Amini, “Using Sliding Meshes”.
208

1. Create periodic zones.


2. Set up the transient solver and cell zone and boundary conditions for a sliding mesh.
3. Set up the mesh interfaces for a periodic sliding mesh model.
4. Sample the time-dependent data and view the mean value.
The methodology is based on the use of Moving Least Squares (MLS) approximation in a high-order
finite volume framework257. Here we present two different approaches based on MLS approximation
for the transmission of information from one grid to another. The intersection approach: the flux at
the interface edge is split between the cell having an interface edge coincident (Figure 9.17 A). The
halo cell approach: a halo-cell is created as a specular image of the interface cell (Figure 9.17 B).
Moreover, two kind of stencil has been tested: the half stencil which take in account only cells from
the grid in which the cell is placed and the full stencil which includes cells from the two grids258.

Figure 9.17 Half Stencil and Full Stencil Reconstruction with: A) Intersection, B) Halo-Cell

9.5.2 Non-Linear Harmonic Balanced Method (NLHB)


The sliding mesh method simulates the full unsteady flow, which is still quite computational
expensive for industrial requirements. In the past decade, a harmonic frequency-domain methods
are developed, e.g., using potential flow model and Euler equations. However, all of the previous
harmonic methods adopt the linear assumption, so that the nonlinear interaction between unsteady
disturbances and the time-averaged flow is completely neglected. A nonlinear harmonic method is
developed by He259 following the framework of Giles260 which is based on an asymptotic theory. In
this technique convergence of Fourier-based time methods applied to turbomachinery flows. The
focus is on the harmonic balance method, which is a time-domain Fourier-based approach
standing as an efficient alternative to classical time marching schemes for periodic flows.
Fourier series decomposes a periodic signal into a sum of an infinite number of harmonics (sine and
cosine functions) of different frequencies and amplitudes. These frequencies are discrete, not all
frequencies are present. Since it is impossible to estimate an infinite series, you choose the number

257 S. Khelladi, X. Nogueira, F. Bakir and I. Colominas, “Toward a higher-order unsteady finite volume solver Based

on reproducing kernel particle method”, Computer Methods in Applied Mechanics and Engineering, 2011.
258 Hongsik, Xiangying Chen, Gecheng Zha, “Simulation of 3D Multistage Axial Compressor Using a Fully

Conservative Sliding Boundary Condition”, multistage turbomachinery are developed and implemented;
Proceedings of the ASME, 2011 International Mechanical Engineering Congress & Exposition IMECE2011,
November 11-17, 2011, Denver, Colorado, USA.
259 L. He, “Modelling issues for computation on unsteady turbomachinery flows. In Unsteady Flows in

Turbomachines”, Von K´arm´an Institute for Fluid Dynamics, 1996.


260 M. B. Giles, “An approach for multi-stage calculations incorporating unsteadiness”, ASME-GT92, number 282,

Cologne, Germany, 1992.


209

of terms you wish to consider, starting from the first. More the number of terms considered, closer is
the series to the original signal. In the literature, no consensus exists concerning the number of
harmonics needed to achieve convergence for turbomachinery stage configurations. It is shown that
the convergence of Fourier-based methods is closely related to the impulsive nature of the flow
solution, which in turbomachines is essentially governed by the characteristics of the passing wakes
between adjacent rows. As a result of the proposed analysis, a priori estimates are provided for the
minimum number of harmonics required to accurately compute a given turbomachinery
configuration. Their application to several contra-rotating open-rotor configurations is assessed,
demonstrating the practical interest of the proposed methodology. This method solves the steady
transport equations for the time-averaged flow and the time harmonics. For turbomachinery, the
Blade Passing Frequencies (BPF) are the fundamentals in time domain of the periodic disturbances
from the adjacent blade rows. The solving of the generated perturbation amplitudes in a row is
performed in the
frequency domain by a
steady transport
equation associated
with BPFs and
subharmonics. The
deterministic stresses
are calculated directly
from the in-phase and
out-of-phase
components of the
solved harmonics.
Using this method, only
one passage is needed
that saves the
computational cost
Figure 9.18 Relative Velocities Obtained using HB Techniques
greatly. He et al., Vilmin
et al. validated this
method with simulations on a 3D radial turbine and a multistage axial compressor. Therefore, this
method is adopted in the unsteady simulation of a low speed axial turbine. The physical quantity can
be decomposed into a time-averaged value and a sum of perturbations, which in turn can be
decomposed into N harmonics261. Figure 9.18 displays Harmonic function method in obtaining
relative velocities (courtesy of NUMECA.com). Since the Harmonic method is widely used, it is
warranted a bit more exploring which will be dealt in the coming section.
9.5.3 Profile Transformation (Pitch Scaling)
In typical turbomachinery applications, it is very common that one or both blade rows have a prime
number of blades per wheel. Formerly in such cases, it was necessary to model the whole 360° wheel
in order to attain the required level of accuracy. It is possible to reduce the size of the computational
problem (memory and computational time) by solving the blade row solution for one or two passages
per row, while still obtaining reasonably accurate solutions, therefore providing a solution to the

S. Vilmin, E. Lorrain, and Ch. Hirsch, ” Unsteady flow modeling across the rotor/stator interface using the
261

non-linear harmonic method”, In ASME-GT06, number 90210, Spain, 2006.


210

unequal pitch problem between the blade


passages of neighboring rows. This
(ANSYS262, Galpin263), a scaling procedure
applied automatically to solution profiles as
part of the TRS implementation, whenever
the rotor-stator pitch ratio is not unity. In
this approximate method, single blade
passages per row with different pitch
lengths can be modeled without the need to
geometrically scale or modify the blade
geometry. Regular periodicity is imposed
for each passage and flow profiles across
rotor/stator interfaces are automatically
stretched or compressed as needed
according to the pitch ratio while Figure 9.19 Phase shifted Periodic Boundary
maintaining full conservation. Multiple
passages can be used to reduce pitch scaling
errors for the ensemble. Since this implementation is fully implicit and conservative a fast and robust
transient solution can be obtained at a fraction of the time for a full domain model. While in this
method overall machine performance is usually predicted well, detailed flow features such as blade
passing signals will be inaccurate due to imposing instantaneous periodicity on the phase-shifted
boundaries264 (see Figure 9.19).
9.5.4 Time Transformation Method (TT) using Phase-Shifted Periodic Boundary Conditions265
Barrowing from ANSYS CFX©, the basic principle of a phase-shifted periodic condition is that the
pitch-wise boundaries R1/R2 and S1/S2 are periodic to each other at different instances in time. For
example the relative position of R1 and S1 at t0 is reproduced between sides R2 and S2 at an earlier
time t0-Δt. Where Δt is defined by (PR-PS)/VR. Here PR and PS are rotor and stator pitches respectively,
and VR is the rotor velocity as shown in Figure 9.20. The Time Transformation method handles the
problem of unequal pitch described above by transforming the time coordinates of the rotor and
stator in the circumferential direction in order to make the models fully periodic in “transformed”
time. Let the r, ϴ, and z coordinate axis represent the radial, tangential (pitch wise) and axial
directions of the problem described in Figure 9.20. Mathematically, the condition of enforcing the
flow spatial periodic boundary conditions on both rotor and stator passages, respectively, is given by

U R1 (r, θ, z, t ) = U R2 (r, θ + PR , z, t − Δt )

 U R1 (r, θ, z, t ) = U R2 (r, θ + PR , z, t)
US1 (r, θ, z, t ) = US2 (r, θ + PS , z, t − Δt ) Eq. 9.5

 US1 (r, θ, z, t ) = US2 (r, θ + PS , z, t )

262 ANSYS CFX Version 12 documentation, ANSYS Inc., 2009.


263 Galpin P.F., Broberg R.B., Hutchinson B.R., “Three-Dimensional Navier Stokes Predictions of Steady State
Rotor/Stator Interaction with Pitch Change”, 3rd Annual Conference of the CFD Society of Canada, June 27-1995,
Banff, Alberta, Canada.
264 “A comparison of advanced numerical techniques to model transient flow in turbomachinery blade rows”,

Proceedings of ASME Turbo Expo 2011 GT2011.


265 ANSYS CFX-Solver Theory Guide, Release 15, 2013.
211

Using the following set of space-time transformations to the problem above as:

Δt
r = r , θ = θ , z = z, t = t − Eq. 9.6
PR − PS

Figure 9.20 Phase Shifted Periodic Boundary Conditions

The equations that are solved are in the computational (r’, ϴ', z’, t’) transformed space-time domain
and need to be transformed back to physical (r, ϴ, z, t) domain before post-processing. The periodicity
is maintained at any instant in time in the computational domain and it is evident that the rotor and
stator passages are marching at different time step sizes. We have the time step sizes
in the rotor and stator related by their pitch ratio as:

PR Δt S
= Eq. 9.7
PS Δt R
Where nΔtS = PR/VR and nΔtR = PS/VR. The simulation time step size set for the run is used in the stator
passage(s) ΔtS and program computes the respective rotor passage time step size ΔtP based on the
rotor-stator interface pitch ratio as described above. When the solution is transformed back to
physical time, the elapsed simulation time is considered the stator simulation time. Required that the
pitch ratio fall within a certain range, as described by the inequality:

Mω P Mω
1−  S  1+ Eq. 9.8
1 − Mθ PR 1 + Mθ

Where Mω is the Mach number associated with the rotor rotational speed (or signal speed in the case
of an inlet disturbance problem), Mϴ is the Mach number associated with the tangential Mach
number, and the ratio of PS to PR is the pitch ratio between the stationary component and the rotating
component. For most compressible turbomachinery applications (for example, gas compressors and
turbines), Mω is in the range of 0.3-0.6, enabling pitch ratios in the range of 0.6-1.5. Note that
according to ANSYS CFX© these limits are not strict, but approaching them can cause solution
212

instability.
9.5.5 Revisiting Non-Linear Harmonic Balance (NLHB) Methodology
Given the time periodic nature of these flows, one can model the unsteady flow in turbomachines
using nonlinear, harmonic balance techniques. Roughly speaking, the family of nonlinear harmonic
methods expands the unsteady flow field in a Fourier series in time and solves for the Fourier
coefficients. [He]266, and [Ning]267 developed a harmonic method in which the unsteady harmonics
are treated as perturbations. [Hall, Thomas, and Clark]268 developed a full harmonic balance method,
which allows for arbitrarily large disturbances and any number of harmonics. The method is
computationally efficient and stores the unsteady nonlinear solutions as the working variables at
several time levels over one period of unsteadiness, rather than storing the Fourier coefficients them-
selves. [Gopinath and Jameson]269 and others have applied this approach to turbomachinery
applications. For an excellent recent survey of Fourier methods applied to turbomachinery
applications, see the survey paper by [He]270. In all these methods, the harmonic balance equations
are solved by introducing a pseudo-time derivative term and then marching the coupled equations
to a steady state. Using the frequency-domain or time-linearized technique, it is possible to first
compute the time-mean (steady) flow by solving the steady flow equations using conventional CFD
techniques. One then assumes that any unsteadiness in the flow is small and harmonic in time (eiωt).
The governing fluid equations of motion and the associated boundary conditions are then linearized
about the mean flow solution to arrive at a set of linear variable coefficients equations that describe
the small disturbance flow. The time derivatives d/dt are replaced by jω where ω is the frequency of
the unsteady disturbance, so that time does not appear explicitly. The resulting time-linearized
equations can be solved very inexpensively, but unfortunately cannot model dynamic nonlinearities.
9.5.5.1 Temporal & Spatial Periodicity Requirement
Consider unsteady flows that are temporally and spatially periodic. In particular, temporal and
spatial periodicity requires that

U (x , t) = U (x , t + T)
Eq. 9.9
U (x + G , t) = U (x , t + Δt)

Where T is the temporal period of the unsteadiness, G is the blade-to-blade gap and Δt is the time lag
associated with the inter blade phase lag. Similarly, for cascade flow problems arising from vibration
of the airfoils with fixed inter blade phase angles σ, or incident gusts that are spatially periodic. As an
example, consider a cascade of airfoils where the source of aerodynamic excitation is blade vibration
with a prescribed inter blade phase angle σ and frequency ω. Then T = 2π/ω and Δt = σ/ω. Because
the flow is temporally periodic, the flow variables may be represented as a Fourier series in time with
spatially varying coefficients.
9.5.5.2 Boundary Conditions
We first consider the flow field kinematics of two adjacent blade rows where the first row has B1

266 He, L., 1996. “Modelling issues for time-marching calculations of unsteady flows, blade row Interaction and
blade flutter”, VKI Lecture Series “Unsteady Flows in Turbomachines”, von Karman Institute for Fluid Dynamics.
267 Ning, W., and He, L., 1998. “Computation of Unsteady Flows around Oscillating Blades Using Linear and Non-

Linear Harmonic Euler Methods”. Journal of Turbomachinery, 120(3), pp. 508–514.


268 Hall, K. C., Thomas, J. P., and Clark, W. S., 2002. “Computation of Unsteady Nonlinear Flows in Cascades Using

a Harmonic Balance Technique”. AIAA Journal, 40(5), May, pp. 879–886.


269 Gopinath, A., and Jameson, A., 2005. “Time Spectral Method for Periodic Unsteady Computations over Two and

Three- Dimensional Bodies”. AIAA Paper 2005-126.


270 He, L., 2010. “Fourier methods for turbomachinery applications”. Progress in Aerospace Sciences.
213

blades spinning with rotational rate ω1 rad/s and the second has B2 blades spinning with rotational
rate ω2 rad/s. The flow field within the stage can be decomposed into a Fourier series in the
rotational direction characterized by a set of Nm1, m2 nodal diameters as

Nm1,m2 = m1B1 + m2 B2 Eq. 9.10

Where m1 and m2 can take on all integer values. In the frame of reference of the first and second
blade row, the frequency of the unsteady disturbance associated with any nodal diameter is

ω1,m2 = m2B2 (ω1 − ω2 ) , ω2,m1 = m1B1 (ω2 − ω1 ) Eq. 9.11

Note that in either row the unsteady frequency associated with a given nodal diameter is a function
of the blade count and relative rotation rate of the adjacent row. Furthermore, associated with each
unsteady frequency is an inter blade phase angle:

B2 B1
σ1,m2 = m2 2π , σ m1 ,2 = m1 2π Eq. 9.12
B1 B2

In the frame of reference of the second row. Clearly the inter blade phase angles associated with a
given nodal diameter are a function of the pitch ratios between the two rows. Note that the pitch in
each row is given by G1 = 2π/B1 and G2 = 2π/B2 in the first and second rows, respectively. Solution
Method Since the solution U is periodic in time, we can represent it by the Fourier series:

M
U ( x, t) =  Uˆ
m=− M
m ( x) e imt
Eq. 9.13
1 N −1 ~
where ˆ
U m ( x) = 
N n =0
U n ( x, t n ) e -imt n

can considered complex conjugate of each other. Here, ω is the fundamental frequency of the
disturbance, M is the number of harmonics retained in the solution: Û m are the Fourier coefficients,
and Ũ n are a set of N = 2M + 1 solutions at discrete time levels tn = nT/N distributed throughout one
period of unsteadiness, T. At any U is vector of conserved variables and can be expressed as

ρ (x, t) =  R n (x, t) eint , ρu (x, t) =  U n (x, t) eint , v( x, t) =  Vn (x, t) eint ,.....
n n n
Eq. 9.14
At any location in the flow field domain we can transform the time level solutions into Fourier
coefficients and vice versa using a discrete Fourier transform operator [E] and its corresponding
inverse E-1 as follows
~ ~
Uˆ = E U or ˆ
U = E−1 U
Eq. 9.15
Where E and E−1 are square matrices of dimension N × N, and the Fourier coefficients and time level
solutions have been assembled into the vectors Ũ as
214

~ ~ ~ ~ ~
U = [U 0 , U1 , U 2 , ....... U N−1 ]T Eq. 9.16

The solutions at each discrete time level are obtained by applying the governing equations to all the
Ũ simultaneously
̃
∂𝐔
∫ ⃗⃗]. d𝐀
dV + ∮[𝐅⃗ − 𝐆 ⃗⃗ = ∫ 𝐒̃dV
V ∂t V
Eq. 9.17
Where the flux and source vectors F͂͂,͂͂ G̃ , and S͂͂͂͂ are evaluated using the corresponding time level
solution. The time derivative in Eq. 9.19 is evaluated by differentiating Eq. 9.15 with respect to time
as follows:
~
U E −1 ˆ E −1 ~ ~
= U= U = [ D] U
t t t Eq. 9.18

Where [D] is the pseudo-spectral, N × N matrix operator. Substituting appropriately, yields the
desired harmonic balance equations:

⃗⃗]. 𝐝𝐀
̃ dV + ∮[𝐅⃗ − 𝐆
∫ [𝐃]𝐔 ⃗⃗ = ∫ 𝐒̃dV
V V
Eq. 9.19
The harmonic balance equations are discretized using a cell centered, polyhedral-based, finite-
volume scheme. Second order spatial accuracy is achieved by means of a multi-dimensional, linear
reconstruction of the solution variables. The convective fluxes are evaluated by a standard upwind,
flux-difference splitting and the diffusive fluxes by a second-order central difference. A pseudo-time
derivative of primitive quantities, ∂Q/∂τ, with Q = {p, u, T}, is introduced into Eq. 9.19 to facilitate
solution of the steady harmonic balance equations by means of a time marching procedure. An Euler
implicit discretization in pseudo-time271 produces the following linearized system of equations:

 U  S U 
 + Δτ [ A] − + [D]  ΔQ = −Δτ R Eq. 9.20
 Q  Q Q 

where R' is the discrete residual of Eqn. (18.17), and ΔQ' are the resultant primitive variable
corrections across one pseudo-time step, Δτ. Operator [A] is the Jacobian of the discrete inviscid and
viscous flux vectors with respect to primitive variables Q and introduces both center coefficients as
well as off-diagonals arising from the linearization of the spatially discretized fluxes. The coupled
system given by Eqn. (18.17) contains equations from all time levels linked at every point in the
domain by the pseudo-spectral operator [D]. The result is a large system, and solving it all at once
would be rather intractable. However, we can exploit the point coupled nature of the system and
employ approximate factorization to produce the following two step scheme:

271Weiss, J. M., Maruszewski, J. P., and Smith, W. A., 1999, “Implicit Solution of Preconditioned Navier-Stokes
Equations Using Algebraic Multigrid”, AIAA Journal, 37(1), Jan., pp. 29–364
215

 U  S  ~
 + Δτ [ A] −  ΔQ = −Δτ R
 Q  Q 
Eq. 9.21
 U −1 U  ~
[ I ] + Δτ [D]  ΔQ = ΔQ
 Q Q 

Where ΔQ̃ ' represents provisional corrections to the solution. In the first step, the time levels are no
longer coupled and we can solve for the ΔQ̃ ' one time level at a time. With the exception of the physical
time derivative appearing, the evaluation of fluxes, accumulation of the residual, and the process of
assembling and solving at each time level proceeds exactly as for a single, steady-state solution in the
time domain. Here we employ an algebraic multigrid (AMG) method to solve the linear system and
obtain the provisional ΔQ̃ ’. In the second step the complete corrections ΔQ' for the current iteration
are obtained by inverting at each point in the domain given all the ΔQ̃ ' computed in step one.
9.5.5.3 Fourier 'Shape Correction' for Single Passage Time-Marching Solution
The Fourier modelling approach to nonlinear flows was proposed in 1990 for time-marching
solutions of unsteady turbomachinery flows272. This was the first Fourier method for
turbomachinery. The objective at the time was to enable an unsteady flow solution to be carried out
in a single blade passage domain but without requiring a large amount of computer memory, as in
the Erdos's Direct Store method. The main ingredient is to carry out the temporal Fourier transform
at the ‘periodic boundaries of the single blade passage domain. Then the Fourier harmonics
(temporal shape) are used to correct the corresponding boundaries according to the phase shift
periodicity. The method was then called ‘Shape Correction’. The validity of the single passage Shape-
Correction method can be examined by comparing with the direct multi-passage solution. Figure
9.21 shows Stagnation Pressure contours under inlet distortion for NASA Rotor 67 where the Left
shows whole passage annulus solution, and the Right, single passage solution as reconstructed. It was
shown that the Fourier modelling as implemented in the Shape-Correction can capture flow
disturbances and responses with large nonlinearity (e.g. a large scale shock oscillation in fan blade
passage under an inlet distortion of long circumferential wave length. Given only 3-5 harmonics were

Figure 9.21 Stagnation Pressure Contours under inlet distortion for NASA Rotor 67

272L. He, "An Euler Solution for Unsteady Flows around Oscillating Blades", ASME, Journal of Turbomachinery,
Vol.112, No.4, pp.714-722, 1990.
216

required for capturing sufficiently accurately the temporal variation, the computer memory
requirement is very low compared to the Erdos’s Direct Store approach. A key advantage of splitting
flow components represented by Fourier harmonics is the ability in dealing with multiple
disturbances with distinctive frequencies (He 1992). The generalized shape correction has been
applied to unsteady flows in multi-
rows (IGT-rotor-stator) with vibrating
rotor blades for optimization of intra-
row gap effects on both aerothermal
performance and flutter stability.
9.5.5.4 Case Study 1 – 2D
Compressor Stage
In this section we compare results
obtained from the implicitly coupled,
non-linear harmonic balance method
described above with solutions from a
full, unsteady simulation based on the
standard dual time-stepping approach.
The test case consists of a model 2D
compressor stage; specifically, the first
stator and second rotor rows of the five Figure 9.22 Computational Mesh for HB and TRS Methods
row. There are three stator blades to
every four rotor blades. The two blade
rows are separated by an axial gap equal
to 0.25 times the aerodynamic chord of
the rotor. The Mach number at the inlet
to the stator is 0.68 and the relative
Mach number entering the rotor is 0.71.
The static-to-total pressure ratio across
the stage is 1.2. Three separate Euler
calculations are made using the
nonlinear harmonic balance method in
which one, two and three harmonics,
respectively, are retained for the blade
passing frequencies in both the stator
and rotor. Contours of instantaneous
pressure, representative of the flow field
within the compressor stage and
computed using three harmonics in each
blade row, are shown in Figure 9.23
using nonlinear harmonic balance
method. Note that computations are
performed on just the center blade
passage outlined in each row. The
solutions shown in the passages above
Figure 9.23 Instantaneous Pressure Distribution Within
and below are phase-shifted the Compressor Stage Using (NLHB)
reconstructions included for clarity.
9.5.5.5 Case Study 2 - 3D Flow in Turbine Cascade
217

3D flow in turbine cascade in which the Harmonic Balance (HB) method is applied for modeling
rotor/stator interaction and pressure fluctuations near trailing edges. Computational results are
compared with Transient Rotor/Stator (TRS) results which shows importance of unsteady effects273.
The harmonic balance method requires only a single blade passage be meshed. A structured HOH
mesh is generated for each of the two blade rows, as shown in Figure 9.22. The inlet and exit grid
planes for each of the blade rows correspond to the axial planes where test data is available. The
blade passage mesh is made up of 1.3 M cells with a near wall spacing of y+ = 1.0 - 2. The HB solver
models the fluid as an ideal gas with turbulence closure provided by the Spalart-Allmaras turbulence
model. The solver is run with a CFL number of 5.0, and separate trials are conducted retaining one,
three, and five modes. The solver has converted to a periodic, unsteady solution within 4000 - 5000
iterations. The TRS solver uses a time step is equal 510-5s. This value correspond 5 steps per vane
passing (10 inner iterations per time step). Assessment of the effectiveness of the two methods of
calculation is carried out on the basis of the comparison of time required for obtaining of non-
stationary periodic solutions on the interval of time, sufficient for the passage of at least one rotation
of the impeller. The HB-results showed that CPU time is increased in 7 times for 5 modes compared
time when calculating with one mode. Using 3 modes CPU time is increased (for one iteration) in 3.3
times in comparison with one-mode approximation. The acceleration of the calculation, which is
defined as the ratio of the CPU time required to obtain a periodic solution using the TRS method to
the CPU time of the decision on the HB method is 1:2 - five modes, 1:1 - three modes and 3:1 - one
mode. Hence, the substantial savings (three times) is observed only in the case of one – mode
approximation. It is important to note that in all cases the calculations were carried out at the same
calculation grid, including two blades. Figure 9.24 present instantaneous predictions of turbulent
viscosity at mid - span for the HB and TRS solutions. The stator wake enters the rotor passage and
grows both laterally and in the stream wise direction. This process continues as the stator wake is
“chopped” by the leading edge of the rotor blade and convects downstream.

TRS HB

Figure 9.24 Instantaneous Predictions of Turbulent Viscosity at Mid-Span for HB and TRS Solutions

9.5.6 Assessment of 2D Steady and Unsteady Adjoint Sensitivities for Rotor-Starter Interaction
Adjoint-based CFD turbomachinery optimization has gained increasing popularity over the last years
due to the advantages in dealing with problems characterized by a large number of design variables

273 Grigoriev A.V., Iakunin A.I.,


Kuznechov N.B., Kondratiev V.F., Kortikov N.N., “Application of Harmonic Balance
Method to The Simulation of Unsteady Rotor/Stator Interaction In The Single Stage”, JSC ‘Klimov’, St - Petersburg,
Russia.
218

at affordable computational cost [Rubino et al.]274. Thanks to this, adjoint-based optimization offers
the possibility to fully exploit the ever increasing computational power to accomplish novel and
unconventional turbomachinery design. To date, despite the intrinsically time-varying nature of
turbomachinery flows, adjoint-based methods mostly rely on steady state approaches. However, the
use of unsteady-based design could lead to major steps forward in performance improvement for the
next generation of turbines and compressor, allowing to tackle multidisciplinary problems. Time-
accurate adjoint methods are well-established but their industrial use is very limited, due to the
excessive computational cost of both the direct flow solution and the I/O overhead associated with
the reverse adjoint mode. The harmonic balance (HB) method is a cost-effective alternative to time
accurate adjoint for non-linear time periodic flow problems, thus it is highly attractive for
turbomachinery applications.
The objective here is to perform an assessment between steady and HB-based unsteady design
sensitivities, investigating the impact of unsteady effects on the aerodynamic design of
turbomachinery. Computational cost, memory requirements are considered as comparison terms,
by using both steady and unsteady methods. The steady-state adjoint calculations are performed by
resorting to mixing plane (MP) approach, whereas the unsteady analysis is carried out with a sliding
mesh interface and solved with a harmonic balance (HB) method. A duality preserving approach is
used in order to ensure robust convergence of the adjoint equations without any restrictive
assumption on the turbulence viscosity. The two methods are implemented in the open-source SU2
software (Palacios et al., 2013; Economon et al., 2015), whose adjoint has been extended in this work
to compute multi-row HB-based sensitivities. The investigation is performed on an axial turbine
stage for both subsonic and transonic conditions, thus resembling the typical flow characteristics of
gas turbine stages. Further information regarding the method of solution can be obtained at [Rubino
et al.]275.
9.5.6.1 Case Study
The test case considered for the present study is a 2D axial turbine stage, adapted from the 1.5 stage
experimental setup of the Institute of Jet Propulsion and Turbomachinery at RWTH Aachen,
Germany276. The mid-span geometries of the first two blade rows, from the above mentioned setup,
are selected for the subsequent analysis reported in this paper. In order to compare the design
sensitivities, for both unsteady and steady state adjoint computations, the proposed test case is
simulated under subsonic (Case1)
and transonic (Case2) conditions, Parameter Case 1 Case 2
thus resembling the typical flow Stator inlet blade angle 0 0
characteristics of a gas turbine stage. Total temperature 305.8 305.8
The main simulation parameters are Pressure ratio 1.5 1.9
reported in Table 9.3. The Rotational speed 3210 4258
simulations are performed using the Inlet turbulence intensity 5% 5%
Roe scheme for the discretization of
the convective fluxes; second order
Table 9.3 Axial turbine simulation parameters
accuracy is achieved by MUSCL
reconstruction. For both unsteady

274A. Rubino, S. Vitale, M. Pini and P. Colonna, “Assessment of fully-turbulent steady and unsteady adjoint
sensitivities for stator-rotor interaction in turbomachinery”, GPPS-NA-2018-130.

275 A. Rubino, S. Vitale, M. Pini and P. Colonna, “Assessment of fully-turbulent steady and unsteady adjoint
sensitivities for stator-rotor interaction in turbomachinery”, GPPS-NA-2018-130.
276 Stephan, B., Gallus, H., and Niehuis, R.. “Experimental investigations of tip clearance flow and its influence on

secondary flows in a 1-1/2 stage axial turbine”, ASME Turbo Expo 2000: Power for Land, Sea, and Air, pages
V001T03A099--V001T03A099. American Society of Mechanical Engineers, 2000.
219

and steady simulations, non-reflective boundary conditions are imposed277 at the stator inlet and at
the rotor outlet sections. The stator-rotor interface is resolved for the unsteady simulations using a
sliding mesh approach, whereas the steady simulations are based on a conservative mixing-plane
(MP) method278. The κ-ω SST turbulent model is considered with fully resolution of the viscous
sublayer. An unstructured grid is used to discretize the 2D computational domain with about 30000
triangular elements for each blade row and 10000 quad elements over each blade surface in order to
ensure y+ ≈ 1. In this work, the selected objective function (OF) for the calculation of the design
sensitivities is the non-dimensional entropy generation of the stage, defined as

⟨Ss,out ⟩ − ⟨Ss,in ⟩ ⟨Sr,out ⟩ − ⟨Sr,in ⟩


Sgen = +
v02 /T0s,in v02 /T0s,in
Eq. 9.22
where ⟨Ss;in⟩ and ⟨Ss;out⟩ are the stator inlet and outlet entropy values averaged over the boundary
using a mixed-out procedure279, whereas ⟨Sr;in⟩ and ⟨Sr,out⟩ indicate the same quantities calculated for
the rotor. v0 is the 'spouting' velocity, namely the velocity that the flow would reach by expanding the
flow isentropically from the total inlet pressure to the stage outlet static pressure.
9.5.6.2 Results
The values of the entropy generation, i.e. the optimization objective function, are calculated for both
mixing plane and harmonic balance simulations of the selected case study. This is accomplished in
order to compare the performance of the stage obtained by the two methods and select an
appropriate number of time instances to resolve. Figure 9.25 shows the evolution in time of the
entropy generation, Sgen, calculated for a different number of time instances as well as the value given
by the steady state mixing-plane (MP) method.

(a) Case 1 - Subsonic (b) Case 2 - Transoinc

Figure 9.25 Non-Dimensional Entropy Generation Using Unsteady (HB) vs Steady (MP)

277 Giles, M. B.. “Nonreflecting boundary conditions for euler equation calculations”. AIAA journal, 1990.
278 Giles, M.. ”A numerical method for the calculation of unsteady flow in turbomachinery”. Technical report,
Cambridge, Mass.: Gas Turbine Laboratory, Massachusetts Institute of Technology, 1991.
279 Saxer, A. P. “A numerical analysis of 3-D inviscid stator/rotor interactions using non-reflecting boundary

conditions”. Technical report, Cambridge, Mass.: Gas Turbine Laboratory, Massachusetts Institute of
Technology, 1992.
220

9.5.6.2.1 Case 1 - Subsonic Stage


Case1 refers to the stage characterized by subsonic conditions. The stage operating conditions
adopted in the simulation are reported in Table 9.3. It was also revealed at [Rubino et al.]280 the
normalized distribution of the static pressure over the blade profiles at different time instances as
well as the time-average HB and the MP solutions. The mixing-plane interface leads to a static
pressure at the outlet of the stator about 4.5% higher than the one attained by the harmonic-balance
simulation, resulting in a lower stator loading and a higher rotor expansion ratio. The steady state
value of Sgen differs of about 4% when compared with the HB time-average solution. The max peak is
about 41% of the mean value, for the entropy generation Figure 9.25 (a), and about 3% in the case
of the stage total-to-static efficiency.
9.5.6.2.2 Case 2 - Transonic Stage
The transonic flow characteristics of Case2 are attained with the same stage geometry of Case1 but
at an expansion ratio about 27% higher (see Table 9.3). The entropy generation predicted by the
MP is about 7.5% lower than the HB time-average value, whereas the relative difference on the total-

(a) MP Pressure Contour (b) HB t = 0

(c) HB t = 1/3 T (d) HB t = 2/3 T

Figure 9.26 Case2: Non-Dimensional Pressure Contours for the Mixing Plane (MP) simulation (a) ,
and Harmonic Balance (HB) at Different Time Instances (bcd)

280A. Rubino, S. Vitale, M. Pini and P. Colonna, “Assessment of fully-turbulent steady and unsteady adjoint
sensitivities for stator-rotor interaction in turbomachinery”, GPPS-NA-2018-130.
221

to-static efficiency is only about 0:3%. As opposed to Case1, in this test case shocks occur on both
stator and rotor. The harmonic-balance simulation proved to be able to reproduce the non-linear
flow characteristics associated to the different stator-rotor mutual positions in time. When
accounting for unsteady effects, both stator and rotor exhibit a stronger shock wave intensity
compared to the mixing plane simulation, with the flow discontinuity appearing at a location different
from that provided by the time-average HB solution. Figure 9.26 reports the non-dimensional
pressure contour plot, for both MP and HB simulations. Further details can be obtained from281.
9.5.6.3 Design Sensitivity
As mentioned in the CASE STUDY section, the entropy generation Sgen(U,a) defined by Eq. 9.22 is
selected as objective function for the optimization problem. A set of design variables corresponding
to the control points of a Free-Form Deformation (FFD) box encapsulating the blade profiles, as
shown in [Rubino et al.]282 for a set of twelve variables on the stator surface. For both mixing-plane
and harmonic balance method, the design sensitivities are retrieved from the adjoint solution
according to
N−1
dO ∂O ∂Fn
= + ∑ λTn
dα ∂α ∂αn
n=0
Eq. 9.23
Where O is the objective function corresponding to α as design variables vector, F is the fixed point
iteration operator, and λn is the adjoint solution. This sensitivities correspond to the gradient given
by the total derivative of the entropy generation with respect to the FFD control points, dsgen /da . The
validation between the objective function gradients obtained with the reverse mode of Algorithmic
Differentiation (AD) and the gradients calculated with second-order finite differences (FD), for Case1
is obtained283. The AD and FD gradients, calculated with respect to a representative ensemble of 24
FFD control points enclosing the stage blade rows, are well in agreement. The Root Mean Square
Error (RMSE) lower than 0:004 for both Mixing Plane and Harmonic Balance method. The same level
of accuracy was achieved when validating AD vs FD gradients for Case2. The computational cost and
memory requirements associated with the HB-based adjoint sensitivities are about 2K+1 higher than
the MP-based gradients computations. The CPU time linearly increases with the number of resolved
number of frequencies K because, for the time domain HB method adopted in this work, the system
of equations is solved in a segregated manner for each time instance. From the details about the
derivation of the HB operator it can be deduced that, for K input frequencies, 2K +1 time instances
must be resolved. Here an odd formulation of the time domain HB method is adopted to preserve
numerical stability284. Since, in terms of CPU time, the mixing plane computation can be regarded
approximately as a single time instance resolution, the associated computational cost is 2K+1 lower
when compared to the HB-based method. The memory requirements follow the same considerations
as for the CPU time: for K resolved harmonics, 2K+1 computational domains, relative to each of the
associated 2K+1 time instances, must be considered. As a result, the memory burden attained by the
HB method is 2K+1 higher than the steady-state calculation.

281 See Previous.


282 A. Rubino, S. Vitale, M. Pini and P. Colonna, “Assessment of fully-turbulent steady and unsteady adjoint
sensitivities for stator-rotor interaction in turbomachinery”, GPPS-NA-2018-130.
283 See Previous.
284 Gopinath, A., Van Der Weide, E., Alonso, J., Jameson, A., Ekici, K., and Hall, K. “Three-dimensional unsteady

multi-stage turbomachinery simulations using the harmonic balance technique”. In 45th AIAA Aerospace
Sciences Meeting and Exhibit, page 892, 2007.
222

With the aim of preliminary assess whether accounting for unsteady effects can influence the optimal
design in stator rotor interaction problems, the adjoint-based design gradients obtained with the MP
and HB method are computed and analyzed for the two mentioned stage operating conditions, i.e.
Case1 and Case2 (see Table 9.3). Furthermore, in order to identify where the main differences
between MP and HB-based sensitivities occur on the computational domain, the absolute value of the
relative difference dsgen is introduced and defined as

dsgen dsgen
( ) −( )
dα HB dα MP
δsgen =
dsgen
( )
⌊ dα HB ⌋
Eq. 9.24
The values of δsgen are computed for each control point of the FFD box and interpolated, for the rest
of the domain, using a bi-cubic
polynomial response surface.
9.5.6.3.1 Case1 - Subsonic Stage
Figure 9.27 depicts the relative
differences between the MP and the
HB objective function gradients,
relative to the set of 24 design
variables. For the stator blade, this
difference is below 7% whereas, for
the rotor blade, it is as high as 63%. In
the stator Figure 9.27(a), the portion
of the domain close to the trailing edge
is the one showing the highest values
of δsgen. This is possibly related to two
main reasons: (a) Case 1 - Stator
• when compared to the sliding-
mesh interpolation used for
the HB method, the MP leads to
a different static pressure at
the stator-rotor interface,
hence at the stator outlet;
• the unsteady potential stator-
rotor interaction effects are
not taken into account by the
MP.

The relative difference between the


MP and the HB gradients is more
remarkable in the rotor, where it is in
average one order of magnitude
higher than in the stator. From Figure
(b) Case 1 - Rotor
9.27 (b) the zone in the proximity of
Figure 9.27 Mixing Plane vs Harmonic Balance Normalized
the rotor leading edge is the one
Entropy Generation Gradients Obtained with the Adjoint
associated with the highest dsgen value. Solution
Such difference is mainly due to the
223

simplification introduced by the MP simulation, in which the stator wake interaction with the rotor
is neglected by resorting to a mixing process at the blade rows interface. Furthermore, at the rotor
inlet boundary, the MP imposes a pressure about 4% higher than that calculated by the HB method.
9.5.6.3.2 Case2 - Transonic Stage
The adjoint-based sensitivities, relative to the test case configuration characterized by a transonic
flow, show an overall outcome comparable with the subsonic stage: the main gradient differences are
associated to the rotor cascade285. Also in this case, the relative difference on the stator are about
one order of magnitude lower than those on the rotor. From a closer inspection, the deviations
between MP and HB-based sensitivities are lower in magnitude when compared to Case1 but,
differently from the subsonic stage of Case1, they also show a sensible contribution by the control
points located near the suction side and the rear part of the rotor blade. This difference can be
explained by recalling that, for Case2, a shock wave pattern crosses the stator-rotor interface. The
resulting flow discontinuity interacts with the stator wake dissipating the associated velocity defects
before reaching the rotor leading edge. However, the shock interacts with the rotor altering its
pressure distribution along the stream-wise direction. The non-linear variation of the shock location,
related to the unsteady position of the rotor in time, is not captured by the mixing plane steady-state
simulation. This results in high values of δsgen not only near the rotor leading edge, as opposed to
Case2.
9.5.6.4 Conclusions
This work documents an assessment between steady and harmonic-balance unsteady adjoint
sensitivities for turbomachinery design problems involving unsteady effects. In this study, the open
source code SU2 was extended in order to deal with unsteady HB multi-row simulations and the
calculation of the corresponding adjoint-based sensitivities. An exemplary axial turbine stage,
operating at subsonic and transonic conditions, was considered. The adjoint-based gradients were
successfully validated against second order finite differences. Results showed that, compared to
steady state calculations, the harmonic balance sensitivities are about 2K +1 more costly, with K the
number of resolved input frequencies. Memory requirements exhibit the same trend, with higher
allocation needed for the harmonic balance computation. Although the steady calculations accurately
predicted the time-average stage performance, the design gradients, as computed by the mixing
plane and the harmonic balance method, were found to be significantly different for both subsonic
and transonic flow conditions. The areas in which this difference was predominant are located in the
proximity of the stator-rotor interface. Possible reasons for such divergence are:
➢ Different pressure imposed by the mixing-plane method at the stator-rotor interface;
➢ The unsteady calculations are able to capture potential and wake-rotor interaction effects.
The assessment conducted in this work indicates that accounting for unsteady effects in the design
process may lead to a different optimal configuration. In order to confirm the present results, current
efforts are devoted to extend the adjoint-based method presented in this study to the shape
optimization of turbomachinery problems involving stator-rotor interactions.

285A. Rubino, S. Vitale, M. Pini and P. Colonna, “Assessment of fully-turbulent steady and unsteady adjoint
sensitivities for stator-rotor interaction in turbomachinery”, GPPS-NA-2018-130.
224

9.6 Case Study - Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design
Perspective286
Citation : Gaetani, P. (2018). Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design
Perspective
The stator-rotor interaction is an important issue in turbomachinery design when the highest
performances are targeted. Different characters mark the interaction process in high-pressure or
low-pressure turbines depending both on the blade height and on the Reynolds number. For small
blade heights, being the stator secondary lows more important, a more complex interaction is found
with respect to the high blades, where the stator blade wake dominates. In low-pressure turbines,
the stator wake promotes the transition to turbulent boundary layer, allowing for an efficient
application of ultra-high lift blades. First, a detailed discussion of the low physics is proposed for
high- and low-pressure turbines. Some of-design conditions are also commented. Then, a design
perspective is given by discussing the effect of the axial gap between the stator and the rotor and by
commenting the effects of three-dimensional design on the interaction.
9.6.1 Introduction
The design of high efficiency axial low turbine stages has to face many challenging problems, and one
of these is connected to the interaction between the stationary and the rotating rows of the machine.
In high-pressure gas turbines, additional issues related to the combustor turbine interaction take
laces leading to further complexity in the design process. The overall context for the design space is,
in fact, an unsteady and three-dimensional low field, where the Mach and the Reynolds numbers vary
along the machine. High-pressure stages typically operate in high-subsonic or transonic regimes and
are normally affected by shock-induced separation on the rotor crown and unsteady stator rear
loading287. Moreover, the high-loading, combined to the low aspect ratio of the first stage blading,
drives the generation of wide swirling structures, whose mixing contributes significantly to the loss
budget288. These secondary lows also affect the low angle distribution and momentum redistribution
inside the blade channel and their accurate prediction is fundamental for the designer of the gas
turbine cooling system.
All of these low structures affect the blade cascade where they are generated and the adjacent ones
in the so-called stator-rotor interaction process. To make clear such a complex low feature, all of
them will be recalled and schematized according to what are available in the open literature. The
primary low structures involved in the interaction process are the wake and the secondary lows.
Many research studies have been proposed in the open literature discussing the wake and the
secondary low evolution and their parametric dependence on the typical turbomachinery
parameters.
The interaction process has been addressed in the last 20 years by many authors both for the high-
pressure stages and for the low-pressure ones. Differences between high- and low-pressure stages
arise for the dependence of the boundary layer and its transition on the Reynolds number. When the
high-pressure stages are of concern, the interaction takes place mainly in terms of shock wave, wake
and secondary lows, leading to the so-called wake-blade and vortex-blade interaction. Thanks to the
high Reynolds number and high inlet turbulence levels, the blade boundary layer state is less
influenced by the incoming viscous structures. It has to be taken into account that also the inlet

286 Paolo Gaetani, “Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design Perspective”, Chapter 5,
http://dx.doi.org/10.5772/intechopen.76009.
287 Denos R, Arts T, Paniagua G, Michelassi V, Martelli F. Investigation of the unsteady rotor aerodynamics in a

transonic turbine stage. ASME Journal of Turbomachinery. 2001;123(1):81-89.


288 Sieverding CH. Recent progress in the understanding of basic aspects of secondary lows in turbine blade

passages. Journal of Engineering for Gas Turbines and Power. 1985;107:248-257.


225

boundary layer properties may cause some pressure fluctuation on the cascade loading, as discussed
in289.
Low-pressure stages, on the contrary, are very sensitive to Reynolds number effects. The wakes
coming from the upstream cascade periodically act as a trigger for the boundary layer transition from
laminar to turbulent conditions. Such periodic transition, possibly re-laminarization, is beneficial in
preventing the boundary layer separation and this allows for higher loading. In this context, ultra-
high lift blade can be proficiently applied either to reduce the aero-engine weight or to power the fan
(among others290-291).
All these issues have been addressed both experimentally and by proper CFD simulations;
experiments require high promptness instrumentation like FRAPP (among others292) or LDV and PIV.
Simulation, as well, requires high performance codes and schemes able to face the sliding of rotors
with respect to the stationary components.
In order to gain a general perspective and to quote the importance of the interaction on the cascade
aerodynamics, the reduced frequency concept has been introduced. It refers to the ratio between the
time scale of the unsteadiness (typically: Ss/U, where Ss is the stator pitch and U is the rotor peripheral
speed) and the one related to the transport of the mass low across the device (i.e. b/Vax where b =
axial chord and Vax = the mean axial velocity component).
The reduced frequency definition then is: f = (bU)/(Ss Vax). When f <<1, the process can be considered
as steady and its time variation related to U can be approximated as a sequence of steady state. When
f >> 1, the process is dominated by the unsteadiness. Finally, when f ≈ 1, the unsteady and quasi steady
processes have the same order of magnitude and importance. In many cases, turbomachinery work
is in the range of f ≈ 1 while for example the combustor-1° stage interaction lies in the quasi steady
conditions293.
In the present contribution, the focus is given mainly to the gas turbines geometries and operating
conditions, even though the same mechanism can be applied to steam stages. As already introduced
by the title, the core of this contribution is devoted to the general discussion of the low physics, rather
than on the quantification and on the detailed description of the specific issues: this way, in author’s
opinion, once the general aspects are acknowledged, the detailed issues; as discussed in papers here
referenced, can be properly understood.
Finally, the discussion will be on a single stage, constituted by a stator and a rotor, taken as a
representative for the whole machine. In the case of multistage turbomachines, the low field
discharge by the rotor will affect, with the same mechanics described in the following, the subsequent
stator. Additionally, there could be some “clocking” features between stators and rotors that may
alter the single stage performance.
Experimental results have been taken by means of a steady five holes probe and fast response
aerodynamic pressure probe (FRAPP) on the high-pressure axial turbine located at the laboratory of
fluid-machinery (LFM) of the Politecnico di Milano. More information on the rig and measurement
techniques reported in various papers. It is important to stress that the FRAPP is applied in a
stationary frame and gives the phase resolved total and static pressure (and hence the Mach number)

289 Hu B, Ouyang H, Jin G-Y, Du Z-H. The influence of the circumferential skew on the unsteady pressure fluctuation

of surfaces of rear stator blades. Journal of Experiments in Fluid Mechanics. 2013.


290 Stieger RD, Hodson HP. The unsteady development of a turbulent wake through a downstream low-pressure

turbine blade passage. Journal of Turbomachinery. 2005;127:388-394


291 Lengani D, Simoni D, Ubaldi M, Zunino P, Bertini F, Michelassi V. Accurate estimation of profile losses and

analysis of loss generation mechanism in turbine cascade. Journal of Turbomachinery. 2017;139:121007/1-9.


DOI: 10.1115/1.4037858
292 Persico G, Gaetani P, Guardone A. Design and analysis of new concept fast-response pressure probes.

Measurement Science and Technology. 2005;16:1741-1750


293 Ong J, Miller RJ. Hot streak and vane coolant migration in a downstream rotor. Journal of Turbomachinery. 7

May 2012;134(5). Article number 051002.


226

and the low angle; then, by assuming a negligible effect of the temperature fluctuations, the relative
Mach number and, by this, the relative total pressure are calculated. CFD results have been obtained
on the same HP turbine geometry by means of Fluent® code.
9.6.2 Stator-Rotor Interaction in Axial Stages
The stator-rotor interaction features different characters if occurred in high-pressure or low-
pressure stages. In low-pressure stages, thanks to the high blade height, the main interaction element
is the wake in a general low-Reynolds environment. On the contrary, high-pressure stages are
typically characterized by small blade heights, due to the high mean density and by a stream with
high Mach and Reynolds numbers and high mean temperatures. As in all stages, the wake generated
by the stator impinges on the rotor blades being an important source of interaction, but due to the
specific features of HP stages–other sources of interaction are present.
The small blade height has the primary impact of powering the effects of the secondary and clearance
lows; in fact, they cannot be considered as negligible and modifies the potential low pattern for a
large amount of the blade span. From the stator-rotor interaction perspective, this feature makes the
problem much more complex as an additional source of interaction takes place.
A common feature of the different kind of secondary lows is to be connected to loss cores, as found
for wakes. However, secondary lows are also vortical structures and hence characterized by vorticity
whose sense of rotation is different among the different vortices. Therefore, in the analysis of the
interaction mechanism, this last feature has to be properly taken into account.
Mach number typically modulates the intensity and position of the swirling cores and, if supersonic,
sets the shock wave pattern discharged by the cascades. The Reynolds number, typically high and for
this the low can be regarded as turbulent, mainly sets the interaction between the incoming
structures and the rotor blades boundary layers. As mentioned earlier, to aid the reader in the
comprehension, the different kinds of interaction are discussed separately.
9.6.2.1 Stator Wake-Rotor Blade
Interaction
The stator wake can be regarded either
as a velocity defect or a loss filament.
According to the first approach, the
velocity triangle composition shows a
very different direction and magnitude
for the relative velocity. Figure 9.28
shows the triangles for the free stream
and for the wake low; it is evident how
the relative velocity of the wake flow
(WW) heads towards the blade suction
side, featuring also a negative
incidence on the rotor blade.
According to the second approach, the
wake has no streamwise vorticity
associated to it, being the only vorticity
present related to the Von Karman
street, whose axis is parallel to the
blade span. Once the wake interacts
with the downstream rotor blade, it is
Figure 9.28 Velocity triangles for the free stream
bowed and then chopped by the rotor (subscript FS) and the wake (subscript W) lows. V =
leading edge. Later on, it is transported absolute velocity, W = relative velocity, U = peripheral
inside the rotor channel, being velocity
smeared and showing two separate
227

legs: one close to the suction side and one to the pressure side. Globally, the wake is pushed towards
the rotor suction side by the cross-passage pressure field and, possibly, its suction side leg may
interact with the blade boundary layer, this feature depends mainly on the rotor loading. Figure
9.29 shows the wake in terms of entropy filament, as computed by CFD in 2D – 1 × 1 case.
Downstream of the rotor blade, the wake typically appears as a distinct loss core close to the rotor
wake or as a part of
the rotor wake; this
option is strictly
dependent also on
the axial position
downstream of the
rotor where the
analysis is done. For
this reason, in some
papers this
mechanism is
acknowledged as
“wake-wake”
interaction.
Being the rotor blade
different in number
with respect to the
stator one, different
rotor channels
experience the
interaction in Figure 9.29 Pattern of entropy evolution (bowing, chopping and transport) of
different time even the stator wake in the rotor channel, as foreseen by CFD
though the basic
mechanism does not
differ. The rate of the interaction depends on the stator wake intensity, that is, on the stator loading,
on the blade trailing edge thickness, on the axial stator-rotor gap, on the Reynolds numbers and, for
cooled blade, on the kind of cooling applied.
In case of low pressure turbines, where typically the Reynolds number is low, as for the aeroengine
cases, the wake – wake interaction is in fact the only effective mechanism. Its importance grows as
the Reynolds number decreases and specifically, the incoming wakes, once interacting with the
suction side boundary layer, promotes the laminar to turbulent transition. Such a transition, on one
hand increases losses but on the other hand increases the boundary layer capability to face adverse
pressure gradient and for this delaying the boundary layer separation and hence the blade stall.
Thanks to this mechanism, aero-engines low-pressure stages have seen an increase of loading and
for this a reduction of weight, either for a reduction of solidity or overall number of rows 294-295. It
has to be recalled that this mechanism constitutes an aerodynamic forcing on the rotor blade whose
frequency depends on the stator blade passing frequency that is in the rotating frame of reference
the frequency which the rotor sees the stator wake passing ahead.
9.6.2.2 Stator Secondary Lows-Rotor Blade Interaction
The basic mechanism for this kind of interaction is the same of the wake-blade one, the vorticial

294 Ravindranah A, Lakshminarayana B. Mean velocity and decay characteristics of the near and far-wake of a
compressor rotor blade of moderate loading. ASME Journal of Engineering for Power. 1980;102(3):535-548
295 Hodson HP, Howell RJ. The role of transition in high lift low pressure turbines. Effects of Aerodynamic

Unsteadiness in Axial Turbomachinery, VKI Lecture Series 2005-03. 2005.


228

filament is bowed, chopped and hence transported in the rotor channel. Notwithstanding such
similarity, two main differences can be acknowledged. The vortical structure has its own streamwise
vorticity in terms of magnitude and sense of rotation and for this a different interaction and impact
with the rotor can be expected depending on the entering position in the rotor channel. Moreover,
the vortex entering in the rotor channel is a low structure specifically localized along the blade span
and pitch, whereas the wake is distributed along the span.
It has to be recalled, without aiming at being exhaustive, that different swirling structures can be
acknowledged downstream of the stator, as depicted in Figure 9.30 and discussed in296. The main
ones are the passage vortices, located symmetrically at tip and hub, activated by the pressure
gradient across the passage and hence directed from the pressure side to the suction side. These
vortices have a wide extension but typically low intensities (i.e. vorticities). At the same time, the
presence of the inlet boundary layer activates also the horseshoe vortices, two legs per end wall.
Coupled to each passage vortex, the shed vortex can be found, activated by the interaction between
the passage vortex and the low momentum fluid belonging to the blade wake. The two passage
vortices have opposite sense of rotation. The two horseshoe vortex legs have opposite sense of
rotation between them and the pressure side leg is co-rotating to the corresponding passage vortex.

Figure 9.30 Simplified Schematic of the Secondary Flows System Downstream of a Rotor

Passage and horseshoe vortices start their growing at the stator leading edge and continue it along
the stator channel, possibly merging among them or smearing depending on the stator loading and
on the inlet boundary layer thickness.
The shed vortex, being activated by the viscous transport, starts growing at the stator trailing edge
at the expense of the passage vortex swirling energy, reaches its highest intensity in about half chord

296Langston LS. Secondary flows in axial turbines; A review. 2006. Annals of the New York Academy of Science.
htps://doi.org/10.1111/j.1749-6632.2001.tb05839.x
229

and then weakens due to the viscous stress that smoothens the velocity gradients. Its sense of
rotation is opposite to the one of the corresponding passage vortex.
Tip clearance vortex may be present depending on the sealing geometries of the stator and of the
rotor. Typically, it is located at the hub in stators while it is at the tip in rotors, this later case being
much more important and frequent. Its sense of rotation is opposite to the passage vortex, being
directed from the pressure side towards the suction side across the blade.
It is important to underline that all these swirling lows are present both in stators and in rotors, but
with opposite sense of rotation as a consequence of the different cross pressure gradient versus in
the two channels.
The secondary flow magnitude and position, besides the difference related to the tip clearance, is
different between the hub and the tip. In fact, the radial equilibrium, that onsets due to the tangential
component downstream of the stator, makes the static pressure at the tip higher than at the hub and
for this a higher Mach number at the hub. The effect of the Mach number is well known and primarily
documented by [Perdichizzi]297. Moreover, the pressure gradient acts to diffuse and to shift
centripetally the vortical structures at the tip and to confine close to the end wall the hub ones.

Figure 9.31 Total pressure loss (Y%), streamwise vorticity (Ωs) and absolute Mach number (M)
downstream of the stator. Experiments at the Fluid machinery Lab. at Politecnico di Milano (Italy).

Possible incidence angles to the stator additionally modulate the secondary lows. Positive incidence
angle strengthens secondary lows, as well as lower solidities, as a consequence of the higher blade

297Perdichizzi A. Mach number effects on secondary low development of a turbine cascade. Journal of
Turbomachinery. 1990;112:643-651. DOI: 10.1115/1.2927705
230

loading. Among others, a global review.


The low entering the rotor is then highly three dimensional and complex, as depicted in Figure 9.31.
In the case presented in Figure 9.31, by the total pressure loss coefficient and the vorticity, the
passage vortices, the shed vortices and a corner vortex can be acknowledged. The Mach number map
is also proposed to show the reduction due to the viscous effects of the wake and vortices and the
modulation by the potential field.
To get the rotor perspective, unsteady measurements performed in the stator-rotor axial gap are
reported in Figure 9.32. Such measurements, taken by FRAPP, have been plotted by applying a
phase-averaging technique and a phase-lag reconstruction. The rotor pitch being smaller than the
stator one, the stator wake in some instant occupies more than half of the rotor pitch, as clearly
evidenced by the total pressure loss map (Figure 9.32(a)). The condition of constant inlet total
pressure both in the absolute and in the relative frame is, so far, an unrealistic condition; Figure
9.32( b) shows the periodic nonuniformities throughout the relative total pressure coefficient
(CPT,R).
To aid the reader comprehension, it is first introduced that, given such complex and stator dependent
low, in this chapter, only the basic low physics is described; in fact, the scope is to provide tools for
the fluid dynamic understanding rather than a unique explanation. As introduced at the beginning of
this paragraph, the basis of the interaction between the vortex filament and the rotor blade field can
be considered as not really different with respect to the wake one. The huge difference consists of the
streamwise vorticity that characterizes the vorticial filament.

Figure 9.32 Rotor inlet low field in the rotating frame of reference. Frame (a) Yloss = total pressure
loss. Frame (b) CPT,R: relative total pressure coefficient. Experiments at the Fluid machinery Lab. at
Politecnico di Milano (Italy).

Once the swirling filament is bended in the rotor channel, the pressure side leg sense of rotation
changes while the suction side leg preserves the original one (Figure 9.33). Moreover, being the
suction side leg accelerated by the overspeed on the rotor section side, its vorticity increases; on the
contrary, the pressure side leg decreases and it is smeared out along its transport.
231

Once the vortical structures enter the rotor channel, they interact both with the passage pressure
field and with the rising vortical structure of
the rotor itself. So far, the stator tip passage
vortices, being opposite to the rotor one, tend
to weaken it (and the same occurs for the hub
ones). On the contrary, the stator shed vortex,
being co-rotating with the rotor passage one
will strengthen it. Swirling lows structure
entering in the rotor close to the end walls will
have stronger effects on the rotor secondary
lows generations; on the contrary, the ones
entering far from end walls will interact in the
downstream portion of the channel.
Moreover, the pressure side legs, as their sense
of rotation is opposite with respect the original
one, will undergo the opposite interaction
features.
As the stator vorticial structures enter the
rotor periodically, with a frequency in the
rotor frame equal to the stator blade passing
one, the interaction process takes places
periodically and this generates a pulsation of
Figure 9.33 Schematics of the stator vortical
the rotor field.
structure transport in the rotor passage
Before discussing in detail the different time
frames, it is straightforward to consider first
the mean flow (Figure 9.34 (a–b)): the CPT,R coefficient is in fact the total pressure in the relative
frame and this evidence the loss cores generated in the rotor and in the stator. The wide low CPT,R
region is mainly due to the rotor wake with some strengthening and enlargement due to the rotor
secondary vortices (tip clearance and tip/hub passage vortex). The vorticial structures can be
acknowledged by making use of the Rankine vortex model applied to the deviation angle map and

Figure 9.34 Time mean flow field downstream of the rotor for a subsonic operating condition (expansion
ratio 1.4, reaction degree at midspan 0.3 and incidence angle close to zero). Frame (a) relative total
pressure coefficient (CPT,R); frame (b) deviation angle (δ). Experiments at the Fluid machinery Lab. at
Politecnico di Milano (Italy).
232

reported in Figure 9.34 b. The clearance low experiences a positive deviation angle as it is less
deflected by the blade than the main low. At the same time, the cross low activated at the hub by the
transversal pressure gradient, generates higher low deflection and for this a negative deviation angle
is found. However, the time mean low field differs from the instantaneous one due to the interaction
process.
The full rotor crown has been calculated by applying a phase leg technique to the experimental
results, measured downstream of the rotor for different stator/rotor phases theory one stator pitch.
It is clearly shown in Figure 9.35 how the 25 channels of the turbine rotor experience different low
conditions, each of them different with respect to the time mean one.
The tip region, being dominated by the tip clearance vortex is weakly sensitive to the periodic low
evolution. On the contrary, the midspan/hub region is strongly periodically pulsating. Being the
stator (n° 21) and rotor blades (n° 25) prime numbers and given that the closest periodicity is around
one-thirds, the pattern evidences a periodicity every 120°.

Figure 9.35 Relative total pressure coefficient on the whole rotor crown. Experiments at the Fluid
machinery Lab. At Politecnico di Milano (Italy).

By considering the total pressure unresolved unsteadiness, calculated as the standard deviation of
the total pressure for each phase and position in the measuring plane , the turbulent structure can be
acknowledged: for this it will be considered as the turbulence (Tu). Some of them are rotor
dependent, like the rotor wake, clearance lows and rotor secondary lows; other structures, on the
233

contrary have a clear periodic evolution with some instant where they do not exist.
Figure 9.36 reports different instants of the rotor evolution for three quantities: the relative total
pressure coefficient, the deviation angle and the unresolved unsteadiness. With respect to the time
averaged low reported in Figure 9.34 (for the same operating condition), the relative total pressure
coefficient shows a fluctuating loss region, with the widest extension at t/BPP = 0.83 and the smallest
one at t/BPP = 0.25, mainly in the hub region. For this later time instant, the deviation angle shows
the smallest gradient in the hub region and the unresolved unsteadiness the lowest intensity.

Figure 9.36 Relative total pressure coefficient (CPT,R), deviation angle (δ) and turbulence (Tu) for 4
interaction phases. Experiments at the Fluid machinery Lab. at Politecnico di Milano (Italy).
234

At t/BPP = 0.37, a vortical structure, evidenced by a high deviation angle gradient, appears in the hub
region and magnifies up to t/BPP = 0.83 where its intensity is the largest. Unresolved unsteadiness
and relative total pressure, mark this structure as a loss core periodically impacting on the rotor
channel. The sense of rotation allows accounting this phenomenon as the impact of the stator hub
shed vortex, strong enough at the stator exit, on the rotor hub passage vortex. The tip region on the
contrary experiences an opposite trend with the strongest vortical structure at t/BPP = 0.25, when
likely the higher inlet total pressure is found.
The highest turbulent and loss contents are found in the tip clearance region due to the high
dissipation related to the clearance low. The interaction here briefly described is unfortunately for
the designer case dependent where, as discussed later one, the axial gap and the loading are
important issues.
To get a comprehensive perspective on the importance and on the region where the interaction takes
place, the standard deviation among the different time fames is straightforward and reported in
Figure 9.37. With reference to the relative total pressure coefficient, the wider fluctuations are, as
qualitatively expected by Figure 9.35 and Figure 9.36 in the midspan-hub region. The more
sensitive region at midspan is located on the wake suction side border that is the place where the
stator wake interacts with the rotor one; in that region, the deviation angle evidences coherently
small fluctuations. The tip region is in fact steady as the clearance low dominates over all other
structures. The hub regions, experience both variation on the total pressure coefficient and deviation
angle, being the seat of the vortex/wake and vortex/vortex interaction. With reference to the
deviation angle map (Figure 9.37 b), the deviation angle experiences the highest fluctuation in the
interface between the tip and hub passage vortex; moreover, all along the rotor wake, it fluctuates,
showing the low turning to be highly sensitive to the periodic interaction.

Figure 9.37 Standard deviation of the Cptr and δ for the different time frames. Experiments at the Fluid
machinery Lab. at Politecnico di Milano (Italy).

9.6.2.3 Off-Design Conditions


To improve the analysis, different operating conditions are described here, being different for the
incidence angle and expansion ratio. The effect of the axial gap will be discussed in the following
chapter. As the interaction depends on the rotor loading, all changes in this parameter will affect the
intensity. Specifically, the increase in the rotor loading, by increasing the incidence angle, will
strengthen the interaction leaving unchanged the basic mechanism. In fact, when the rotor loading
increases, the rotor suction side boundary layer is more prone to instability and the rotor vortical
structure more intense, making all of them more sensitive to any variation coming from upstream.
Figure 9.38, shows the relative total pressure coefficient and deviation angle standard deviations,
235

calculated among the different time instants, for a negative incidence conditions (Figure 9.38 a,
incidence at midspan = −10°) and a positive one (Figure 9.38 b, incidence at midspan = +10°). When
the overall effect is of concern, the different interaction intensity leaves a trace on the total to total
efficiency that can be summarized by stating that the higher the interaction, the lower is the
efficiency. When the fluid-dynamic forcing on the rotor blade is under study, the frequency of the
forcing event is the one of the stator passing frequency multiplied by the number of swirling
structures found along the pitch. It has to be brought to the attention of the reader that the stator-
rotor interaction is fundamental for the analysis of the interconnection frames, as discussed in and
of the turbine acoustic behavior298.

Figure 9.38 Rotor loading effects on the stator-rotor interaction. Experiments at the Fluid machinery
Lab. at Politecnico di Milano (Italy).

9.6.2.4 Stator Shock-Rotor Blade Interaction


The third possible source of interaction is related to the shocks generated at the stator trailing edge
that impinge on the rotor leading edge region. Stators in high-pressure stages work often in transonic
conditions at least in the hub region. Rarely, they are chocked as in this condition low rate regulation
is limited. The shock system typically has a fish-tail pattern characterized by oblique shocks; the

298Knobloch K, Holewa A, Guérin S, Mahmoudi Y, Hynes T, Bake F. Noise transmission characteristics of a high
pressure turbine stage. 22nd AIAA/CEAS Aeroacoustics Conference; Lyon, France; 2016.
236

suction side shock is stronger than the pressure side one. The suction side shock propagates
downstream and interacts with the following row, while the pressure side one impinges on the
adjacent blade, specifically on the suction side, being further reflected downstream.
Across the shock, the low experiences a steep and opposite pressure gradient that, if applied to the
boundary layer, acts to de-stabiles it, leading to separated low bubbles. Thanks to the high Reynolds
number, whose action is to promote the momentum exchange in the boundary layer, the effect is not
that critical in high-pressure stages; it has to be recalled that rarely the outlet Mach number exceed
1.5, value where the entropy rise due to shock starts to be important.
As the stator shock sweep the rotor leading edge region, unsteadiness in the static pressure is found
and for this in the boundary layer evolution; luckily, this happens where the boundary layer
momentum deficit is close to be the smallest at the very beginning of the boundary layer evolution.
As reported by [1, 36–38], the rotor trailing edge region is slightly affected, at least in term of static
pressure and for this the boundary layer and the rotor wake are expected to be almost steady. The
highest interaction is found in the leading edge/suction side region as clearly reported in Figure
9.39; the shock sweeping on the rotor leading edge first interact with the suction side of the blade
(approx. in the location of measuring point n° 6, in Figure 9.39) and then reached the leading edge
(measuring point n° 2).

Figure 9.39 Vane shock-rotor interaction in axial turbine blades. Red: computation, black: experiments.
Adapted from [Denos et al.]

The pressure side is less affected by the interaction being overshadowed by the leading edge. It is
clear how the blade shape, in terms of camber/stagger angles and front/rear loading as well, it is a
key parameter for this class of interaction. The magnitude of the stator shock impinging on the rotor
is strictly dependent on the axial gap; the wider it is, the weaker is the shock effect, being the shock
decay rather fast.
From a mechanical perspective, the forcing induced by the stator shock on the rotor is at the stator
passing frequency multiplied by the number of shocks impinging on the rotor per each stator
passage, even though typically only one is important. Other interesting studies on the interaction in
transonic turbine are, where different conditions and geometries are discussed.
237

9.6.3 Design Perspective


In general, there are a huge number of parameters that can be adjusted during the design process.
Among the different parameters, some of them will be hereby described to deepen the understanding
of the interaction features.
9.6.3.1 Axial Gap
The axial gap is one of the key parameter for the stage optimization. In general, the increase of the
axial gap promotes the wake and secondary lows mixing and this leads to a more uniform rotor inlet
low field in the absolute frame of reference. However, the mixing increases losses and the overall
total pressure level reduces. For low axial gaps, on the contrary, low-mixing takes places but a highly
nonuniform low enters in the rotor, leading to additional losses in the rotor itself. It is clear so far,
how the axial gap is a parameter that has to undergo an optimization process and this is the reason
why it has been the focus of a number of research that gave different results, likely depending on the
operating condition and stage loading.
In the context of the wide experimental campaign on the stator-rotor interaction at Politecnico di
Milano, the axial gap has been also addressed and studied. The detailed discussion of the results is
reported in299, while in this context only a brief recall is proposed. Three gaps have been
experimentally investigated, equal to 16, 35 and 50% of the stator axial chord. For the lowest gap
case, the stator structures like wake and passage vortices are more intense than other cases and this
promotes a strong interaction that results in a severe periodic fluctuation in the rotor outlet
quantities. On the contrary as the gap increases, the mechanism is mainly driven by the stator shed
vorticities that strengthen at the expense of the passage vortices intensities, as described in the
previous paragraph. The somehow surprising result is that the lowest interaction rate is for the
design case that is a gap of 35% of the stator axial chord, condition where the inlet low field is the
more uniform in terms of relative total pressure and rotor deviation angle, as clearly depicted in
Figure 9.40. For larger gaps, the combination of stator potential field and wake acts to amplify the
inlet fluctuation, as reported for the incidence angle in Figure 9.41.

Figure 9.40 Standard deviation for the different instants of the interaction phases. (A) axial gap: x/bs =
16%; (B) axial gap: x/bs = 35%, nominal; C) axial gap = 50%. Experiments at the Fluid machinery Lab. at
Politecnico di Milano (Italy).

299Gaetani P, Persico G, Osnaghi C. Effects of axial gap on the vane-rotor interaction in a low aspect ratio turbine
stage. Journal of Propulsion and Power. 2010;26:325-334
238

Overall, the stage experiences the


maximum efficiency for the design case
(Figure 9.42), about 1% higher,
showing the potential of this
parameter in the optimization during
the design process. According to the
open literature, this trend is confirmed
by some authors but seems not
general, either for a lack of detailed
data or for a case dependency in the
stator wake-potential field coupling
along the axial direction, in the axial
gap region.
When the aerodynamic forcing is of
concern, the axial gap plays a role,
since the forcing functions, as the wake
velocity defect and the secondary lows,
are stronger for low axial gaps. Figure 9.41 Rotor incidence fluctuations in
circumferential direction for the different axial gaps along
9.6.3.2 End wall Contouring and 3D the blade span
Blade Geometries
As discussed in the previous sections,
the secondary lows are in the high-pressure
turbine stages a leading issue for the
cascade interactions. Given this matter of
fact, any action devoted to the secondary
low reduction or segregation is
straightforward for softening the stator-
rotor interaction, aiming moreover to an
overall efficiency increase. Among the
possible turbomachinery design
methodologies, two of them are here briefly
commented: the end wall contouring and
the 3D blade design.
The end wall contouring consists of a
specific end wall shape, at tip/hub or both,
aiming at providing lower velocities in the
blade portion where the highest loading is Figure 9.42 Efficiency trend versus the axial gap.
Experiments at the Fluid machinery Lab. at Politecnico
applied, that is higher turning. This feature
di Milano (Italy).
results in a lower local cross-passage
pressure gradient and a strong acceleration
in the rear part of the blade. The final result is a reduction of the passage vortex in the contoured side
of the passage, while the not contoured side experience about the same vortical structures.
The second possible action is the 3D blade design. It consists of a design methodology based on a
different blade stacking with respect to the conventional radial one, leading to the so-called leaned
and/or bowed blades. Typically, the lean given to blades is positive that means a blade stacking
inclined towards the pressure side. The bowing is given by applying a symmetrical leaning at tip and
hub. These methodologies allow for a “low control’ at the cascade outlet in terms of radial pressure
gradient and hence reaction. In case of positive leaning, an additional vorticity is introduced on the
channel; specifically, it increases the one related to the passage vortex at the hub and smear the tip
239

one. At the same time, the lean change the blade loading along the span by amplifying the tip one and
reducing the hub one: overall such a feature makes the secondary low at the hub less intense than
the case of prismatic blades.
Figure 9.31 shows the vorticity field downstream of the lean annular cascade, characterized by a
positive lean of 10°. When the lean is applied symmetrically at hub and tip, this benefit is gained also
at tip. Overall, the final effect in the frame of the stator/rotor interaction process is the reduction of
the secondary lows and their segregation at the end walls. This design methodology leads to an
overall benefit even though the single cascade does not improve significantly its performance.
9.6.3.3 Cascades Clocking
Cascades clocking refers to the design option related to the proper alignment of blades belonging to
different cascades in the same frame of reference (stator/stator or rotor/rotor) in the context of
multistage machines. In fact, downstream of each stage the “wake avenue’ is found, that is the global
effect of the stator wake and secondary lows on the rotor outlet low field in the time mean context.
This concept can be applied also to the rotor wake, when a multistage environment is considered. To
ease the understanding, let us refer to a two stages machine and specifically to the impact of the 1°
stator wake avenue on the 2° stator. Depending on the kind of stages and their loading,
the impact of the wake avenue can be proficiently used for increasing the following cascade efficiency.
In order to clock the two different rows, cascades should have number of blades that are multiple
each other and for this the design assumption of prime number have to be abandoned (it can be kept
for the stator and rotor of the single stage). So far, the highest efficiency is therefore gained by using
the same blade numbers between the two stators (or rotors for rotor-clocking).
In LP turbines the clocking is directly linked to the wakes, while in transonic HP turbines the effect is
mainly related to the interaction of wakes and secondary lows, that is the total pressure and total
temperature fields on the whole, downstream the first stage with the second stator.
According to the early work from300, the efficiency is achieved when the segments of the first vane
wake avenue, released by the rotor, impinge on the leading edge of the second vane. The basic reason
for this result is that the low momentum fluid coming from the first stage, collapse in the boundary
layer of the second vane and for this do not affect the passage, studied in detail the clocking effects
driven by the stator secondary lows in a two stage subsonic and transonic turbines.
According to [Schennach et al.]301, the interaction with secondary vortices is highly complex due to
the different kind and intensity of the vortical structure itself. When the rotor structures dominate,
as can happen in the tip region due to the tip clearance or for the hub secondary vortex, the clocking
effect is somehow shadowed. The outer part of the channel, being typically the rotor tip passage
vortex highly sensitive to the stator-rotor interaction in the upstream stage, is the place with the
highest potential for the clocking. This result makes the proper alignment choice complex for the
designer as it is not really general.
In case of transonic stages, the hub region, being the seat of the 1° stator shock wave and hence of
the highest stator-rotor modulation, has a high potential for clocking.
As a general conclusion, when the low momentum fluid enters on the 2° stator leading edge or close
to the pressure side, the highest efficiency is found. On the contrary, when the low momentum fluid
coming from the 1° vane enters close to the 2° vane suction side, the lowest efficiency is found, as a
consequence of the destabilizing effects on the suction side boundary layer and the lowest expansion
ratio there available and for this lowest suction side overspeed and for this blade lift.
Low-pressure turbines behavior, where an increase of 0.7% in the efficiency is found by numerical
simulations.

300 Jiang JP, Li JW, Cai GB, Wang J. Effects of axial gap on aerodynamic force and response of shrouded and
unshrouded blade. Science China Technological Sciences. 2017;60(4):491-500
301 Schennach O, Woisetschläger J, Paradiso B, Persico G, Gaetani P. Three dimensional clocking effects in a one

and a half stage transonic turbine. Journal of Turbomachinery. 2010;132(1):011019. 1-10. Inglese
240

As a conclusive comment, the benefit achievable by clocking the cascades can be of the order of 1%
in the 2° stator efficiency, being anyway highly depending on the stage features.
241

10 Axial Turbo-Machinery Design and Optimization


10.1 A Road Map to Turbo-Machine Design and Optimization
The design of a turbomachine can be traced through three basic phases. First is the preliminary
design phase in which the type of machine to be employed is determined. Additionally, the size,
speed, and over-all geometry are determined. Since the entire design process is iterative, the
preliminary design parameters and shapes are always subject to modification. Many aspects of this
preliminary design process are
empirical and/or arbitrary and
are based on engineering
experience, system or installation
limitations, costs, and other
factors of which the designer must
be aware. However, to be able to
proceed with a detailed design, a
fairly complete conceptual form of
the machine must be generated.
The second phase is the detailed
design of the machine duct and
blade shapes. A design that is
based on the equations of fluid
motion requires the development
of a mathematical description or
model of the flow field and the
machine geometry. This phase
requires computerized methods
to solve the equations of motion
while accounting for as many of
the physical properties and
boundary conditions as possible.
Figure 10.1 General Description of Computational Planes
Since the direct solution of the
equations of motion for viscous,
turbulent flow is impossible, it becomes necessary to approximate some of the physics of the flow.
However, the exact solution of the no viscous or ideal flow field with forces due to fluid accelerations,
rotation, and non-uniform flows taken into account is possible. Figure 10.1 helps to visualize these
various aspects of the flow geometry. The usual approach used is to solve for the inviscid flow field
and superimpose the effects of real fluid flow which are difficult to treat analytically. The solution
must include physical effects peculiar to each type of machine. Generally, the inflow conditions to the
turbomachine will be no uniform in velocity and pressure. The chord wise and span wise loading or
pressure on the blade rows will be no uniform as is the blockage due to blade thickness and boundary
layer growth. If the flow field is unbounded, as for a nonconductor fan, additional boundary
conditions are necessary to perform a flow analysis. The Effects of viscosity in the blade row passages
must be modeled accurately to achieve the design performance requirements. The second phase
terminates in an actual blade shape design that will produce the flow field specified by the analysis.
Determination of this shape is difficult as there are a number of boundary conditions that must
satisfy, and the performance of a blade row is highly subject to real flow phenomena which are not
easily approached analytically. Again, a combination of ideal flow theory and empirical corrections
are required to specify a blade row with desired performance.
242

The third phase is essentially similar


to the analysis and blade design
except that the geometry of the system
is fixed and it is desired to determine
the effect of a blade row on the flow

Phase 1
field. This should give the same
solution as the original analysis at the
design point. However, the results of
performance testing and, in particular,
measured velocity and pressure
profiles in the vicinity of the blade row
may be used to test the theoretical
analysis and design. Where
differences are found, corrections to
the mathematical models or to the
hardware may be required.
In summary, the design and analysis
of a turbomachine requires an

Phase 2
accurate description of the internal
flow field. A numerical simulation of
the turbomachine, including the
various components of the
equations of motion in either exact
or approximate (empirical) form,
must be constructed. This
simulation consists of two basic
parts, a meridional plane or
through-flow solution and a blade
design method that includes an
analytical blade-to-blade flow

Phase 3
solution. Each of these solutions
affects the other and must be
developed iteratively to produce a
consistent model of the flow. Once a
model is generated, a check against
the desi, “an requirements must be
made to assure that no part of it fails
the design requirements. Once Figure 10.2 Turbomachine Design Process
complete, the model must be
compatible with mechanical limitations and manufacturing capabilities. An additional iteration
with the design constraints may be necessary to finally arrive at an acceptable engineering
solution to the design of the turbomachine. Figure 10.2 illustrated these steps require in
turbomachinery design and analysis302.

302M. V. Casey, “Computational Methods for Preliminary Design and Geometry Definition in Turbomachinery”,
Fluid Dynamics Laboratory, Sulzer Innotec AG, CH-8401, Winterthur, Switzerland.
243

10.2 Optimization Methods for Turbomachinery Designs


According to [Li & Zheng]303, aerodynamic design optimization methods can be distinguished into
inverse and direct designs. Inverse design methods specify a pressure distribution to develop a
profile shape by iterative modifications of the blade shape. The computational cost is proportional to
a small number of flow analyses and is, thus, comparably inexpensive. Inverse design methods can
be combined with an optimization method in an efficient design process. However, the pressure
distribution may pass through a number of iterations to obtain an acceptable profile. This approach
relies strongly on the experience of the designer who needs to specify a pressure distribution which
fulfills the various aerodynamic design aspects in terms of the flow turns, boundary layer properties
and flow losses which also performs well for off-design conditions. Another shortcoming of the
inverse design method is how to integrate the geometric and mechanical constraints. Unlike with
inverse design process, the direct design method optimizes the shape based secondary aerodynamic
properties like the aerodynamic losses with the computational cost involving many single flow
calculations. The optimization algorithms used with the direct design method are mainly the
gradient based methods and the stochastic algorithms. Gradient-based methods rely on derivative
information for all the objectives and all the constraints to determine the optimization search
direction. These methods start with a single design point and use the local gradient of the objective
function with respect to changes in the design variables to determine a search direction by using
methods such as the steepest descent method, conjugate gradient method, quasi-Newton techniques,

Traditional
Gradient Method

Adjoint Variable
(AV)
Gradient Based
Algorithms
Conjugate Gradient
Method

Quasi-Newton
Techniques
Direct Methods

Genetic Algorithms

Turbomachinery
Evolutionary
Optimization Algorithms
Process Stochastic
Algorithms
Simulated
Annealing (SA)
Inverse Methods Set of Equations
Particle Swarm
Optimization (PSO)

Figure 10.3 Optimization Process for Turbomachinery

303Zhihui Li, Xinqian Zheng, “Review of design optimization methods for turbomachinery aerodynamics”,
Progress in Aerospace Sciences, July 2017.
244

or adjoint formulations. These methods are efficient and can find a true optimum as long as the
objective function is differentiable and convex. However, the optimization process can sometimes
lead to a local, not necessarily a global, optimum close to the starting point. Furthermore, such
computations can easily get bogged down when many constraints are considered.
Genetic Algorithms and Evolutionary algorithms are typical stochastic optimization algorithms.
These methods are robust optimization algorithms that can cope with noisy, multimodal functions,
but are also computationally expensive in terms of the necessary number of flow analyses required
for convergence. They start with multiple points sprinkled over the entire design space and search
for true optimums based on the objective function instead of the local gradient information by using
selection, recombination, and mutation operations. Figure 10.3 shows a typical optimization
process for turbomachinery application.
10.2.1 Wu’s Pioneering (S1 and S2) Scheme
In the early 1950's, Wu304, as previously recognized, documented these problems and formulated a
set of equations which had the possibility of a solution. He broke the problem of 3D flow into a set of
coupled 2D solutions. Figure 7.1 helps to explain Wu's analysis in which he broke the problem into
two planes (S1 and S2) generally perpendicular to each other. One, the meridional plane, describes
the flow on hub-to-tip stream surfaces. The other, the blade-to-blade solution, describes the flow on
planes generally parallel to the hub surface of the machine and perpendicular to the blading. A
complete solution by Wu's method would require a number of both parallel meridional and parallel
blade-to-blade solutions. The solutions are coupled and must be solved iteratively to simultaneously
satisfy the equations on all of the solution planes (Quasi-3D). At the time of formulation of Wu's
analysis, computational methods and machines were not large or fast enough to give a
comprehensive solution. As a result, many approximate methods evolved. [Wislicenus]305
summarized many of the design techniques in use at the time. Most of these techniques relied heavily
on experimental data to be useful. Smith 306 rearranged the equations of motion in the meridional
plane to give a time and spatially averaged picture of the flow in a blade row. At the same time,
additional computerized techniques were developed to solve the through-flow problem. [Novak]307
formulated a solution (Streamline Curvature Method) that solved for the velocities and streamlines
rather than the stream functions where solution was basically inviscid and non-turbulent. The
problem of losses due to viscosity and turbulence was addressed by Bosman and Marsh308, but in
general, experimental data are always required to adequately model the real fluid effects
encountered in a turbo machine.
10.2.2 Concept of Streamline Curvature Method
The Streamline Curvature Method (SCM) offers an advantage in that the equations and solution are
in terms of physical variables of velocity and pressure rather than those of a stream function, which
previously mentioned. Additionally, viscous and turbulence effects are much easier to incorporate
into the (SCM) because their models are developed in terms of physical variables. Where the 3D
nature of the flow field is required, determination effects of the blading on the meridional flow
requires flow field solutions on the blade-to-blade surfaces. The design of blade sections also requires

304 Wu, C. H., "A General Theory of Three Dimensional Flow in Subsonic and Supersonic Turbomachines of Axial,
Radial, and Mixed Flow Types," National Advisory Committee on Aeronautics, NACA TN 2604, 1952.
305 Wislicenus, C. F., Fluid Mechanics of Turbomachinery, New York, N. Y., Dover Publications Inc., 1965.
306 Smith, L. H., Jr., "The Radial Equilibrium Equation of Turbomachinery," Trans. ASME, J. Eng. Power, 1966.
307 Novak, R. A., "Streamline Curvature Computing Procedures for Fluid Flow Problems," Trans. ASME, J. Eng.

Power, Vol. 89, 1967.


308 Bosman, C, and H. Marsh, "An Improved Method for Calculating the Flow in Turbomachines, including a

Consistent Loss Model," J. Mech. Eng. Sci., 1974.


245

an accurate analysis technique. [Kansan’s]309 was one first to successfully compute the velocity and
pressure distribution on the blade-to-blade plane. None of the above methods incorporates sufficient
modeling of the turbulent boundary layer flow associated with turbomachine blade rows. [Raj and
Lakshminarayana]310 conducted experiments which gave insight into the nature of the blade
boundary layers and the structure of the wake shed from the trailing edge of a blade. These data will
help in the formulation of more accurate models of this flow phenomenon. The availability of the
various through-flow and blade-to-blade solutions leads to the possibility of synthesizing a three-
dimensional model of the turbomachine flow field. The interaction of the flow on the meridional
plane and on the blade-to-blade planes becomes important in this case. The result of blade-to-blade
analysis is that forces due to the geometry of the blading may be determined. [Novak and Hearsey]311
utilized the Streamline Curvature Method in a similar manner to generate a quasi-three-dimensional
analysis. It should be stressed that the above techniques are for analysis of already designed blade
rows and do not apply to the actual determination of a blade shape, i.e., the design problem. The aim
here to address the design problem by using the Streamline Curvature Method to construct an
averaged through flow picture that satisfies the general design requirements. Then two methods, the
Mean Streamline Method and the Streamline Curvature Method, will be used to actually define
blading that generates the flow field prescribed by the through-flow analysis. Combining the
through-flow and the blade-to-blade analysis, a quasi-three-dimensional analytical
representation of the flow field is generated.

10.3 Case Study 1 - Aerodynamic Design of Compressors


Due to low cost and speed of CFD comparing to traditional testing, and the fact that it ability to
simulated almost any testing, it can be useful tool in design and optimization. While experiment
yields a discrete and limited data, CFD provide the entire region on interest or complete picture
enabling complete design. In that respect, aerodynamic design techniques of gas turbine compressors
have been dramatically changed in the last few years. While the traditional 1D and 2D design
procedures are consolidated for preliminary calculations, emergence techniques have been
developed and are being used almost routinely within industries and academia. The compressor
design still remains a very complex and multidisciplinary task, where aero-thermodynamic issues
traditionally considered prevalent, now become part of a more general design approach. Nowadays,
interesting and alternative options are in fact available for compressor 3D design, such a new blade
shapes for improved efficiency, end-wall contouring and casting treatment for enhanced stall margin,
as well as many others. For this reason while experimental activity remains decisive for ultimate
assessment of design choices, numerical optimization techniques, along CFD are assuming more and
more importance for the detailed design.
10.3.1 Statement of Problem
Qualitatively speaking, compressor aerodynamic design is the procedure by which the compressor
geometry is calculated which fulfils the design cycle requirements in the best possible way312. Using
a more certain statement, we can formulate the design problem by identifying objectives, boundary
conditions, constraints, and decision variables as follows:

309 Katsanis, T., "Use of Arbitrary Quasi-Orthogonal for Calculating Flow Distribution on a Blade-to-Blade
Surface in a Turbomachine." NASA TN D-2809, May 1965.
310 Raj, R., and B. Lakshminarayana, "Characteristics of the Wake Behind a Cascade of Airfoils," J. Fluid

Mechanics, Vol. 61, Part 4, 1973.


311 Novak, R. A., and R. M. Hearsey, "A Nearly Three-Dimensional Intra Blade Computing System for

Turbomachinery," Part I and TI, Trans. ASIT4, J. Fluids Eng., March 1977.
312 Benini, E., “Advances in Aerodynamic Design of Gas Turbines Compressors “, University of Padova Italy.
246

➢ Objectives: Maximize Adiabatic Efficiency (η), Maximize stall margin (SM), both at nominal
condition.
➢ Boundary conditions: inlet conditions (pressure P, temperature T), flight Mach number, M,
(in the case of an aero-engine).
➢ Decision (Design) variables: number of stages, compressor and stage geometry parameters.
➢ Functional constraints: mass flow rate, m, (based on engine Power or Thrust requirements),
Pressure Ratio, PR, (from Cycle analysis), correct compressor component matching (i.e.
intake-compressor, compressor-combustor, and above all compressor-turbine) as
determined by a Matching Index (MI).
➢ Side Constraints: each decision variables must be chosen within feasible lower and upper
bounds (sides).
➢ Multi-disciplinary constraints: structural and vibrational, weight, costs, manufacturability,
accessibility, reliability.

The aerodynamic design of an axial-flow compressor is inherently a Multi-Objective Constrained


Optimization problem, which can be summarized on the as:

X = X (x1 , x 2 ,......, x n ) x i,min  x i  x i,max for i = 1, 2,......n


Eq. 10.1
Where n is the number of decision variable. Table 10.1 summarizes these decision making criteria.
It is worth underlying that this might not be the general formulation, as some constraints could be
turned into objectives, mainly depending on compressor’s final destination and/or manufacturer’s
strategies.

For a Maximize: Minimize: Subject to: Comments:


given:
Axial Flow P ,T , F(X) = η (X) , m(X) , PR(X), g(X) = weight,
Compressor M SM(X) MI(X) , Cost structural,
function (X), technological, other
g(X)
Stationary gas P ,T F(X) = η (X) , m(X) ,PR(X), g(X) = weight,
Turbine Load – MI(X), Cost structural,
(Electric Power) Response (X) function(X) , technological, other
g(X)
Aero-engine P, T, M F(X) = η (X) , weight (X) m(X) , PR(X) g(X) = structural,
(military use) SM(X) , q(X) , MI(X) technological, other
Cost
function(X), q(X) = static trust at
g(X) sea level

Table 10.1 Axial Flow Compressor Design

10.3.2 Different Compressors Objectives


In a stationary gas turbine used for electric power generation a great importance is given both to the
compressor peak efficiency, and to the function called “load response” which quantifies the rapidity
of the compressor in adjusting the airflow delivered by means of IGVs and/or VSVs. In this case, of
course, an intervention aimed at regulating the delivered power of the gas turbine has an effect also
on the mass flow conveyed by the compressor. An aero-engine for military use, a significant merit is
247

attributed to reaching the best trade-off between performance and weight, objectives which are
intuitively conflicting each other, while cost function is inevitably different to the one assigned to the
civil application.
Finally, a constraint based on the static thrust to be delivered by the overall engine at sea level is set
which inevitably influences compressor design. From the problem formulations given above,
remarkable importance is attributed to maximize or minimize some compressor performance
indexes or figures of merit. Therefore, before examining how to deal with such problems, it is worth
analyzing how performance can be significantly affected by the choice in the design variables. For
instance, maximizing adiabatic efficiency requires a deep understanding of the physics governing
stage losses, which have to be minimized either in design and off-design conditions. This, in turn, will
have an important impact on the choice of stage geometrical and functional variables. On the other
hand, maximizing stall margin involves acquiring a proper insight of stall physics and minimizing
stall losses. Again, such problem can be tackled if proper stage geometry is foreseen. Lastly,
minimizing compressor weight (at least from the aerodynamic point of view) implicates reducing the
number of compressor stages and increasing individual stage loading, a fact which ultimately affects
the choice of the blade shape, particularly cascade parameters. Based on the arguments above, in the
following a brief summary of basic and advanced compressor aerodynamics is given313.

Figure 10.4 Sketch of a Compressor Stage (left) and Cascade of Geometries at Mid- Span (right)

10.3.3 Design Techniques for Compressor314


Independently from the particular case under study, modern compressor design philosophy can be
summarized as in Figure 10.5. A preliminary design is usually carried out at first, aiming at defining
some basic features such as number of stages, inlet and outlet radii and length. Stage loading and
reaction is established as well on the basis of preliminary criteria driven by basic theory and
experiments. Such procedure is based on one-dimensional (1D) methods, where each stage
characteristics are condensed into a single “design block”, to which basic thermo- and fluid dynamics
equations are applied. Therefore, no effort is spent to account for flow variations other than those
which characterize the main axial flow direction within each stage at a time. Then, stages are stacked
together to determine the overall compressor design, regardless any mutual stage interaction. Within

313 Benini, E., “Advances in Aerodynamic Design of Gas Turbines Compressors “, University of Padova Italy.
314 Same source – see above.
248

this process, which will be described later on, technological and process constraints, as well as
restrictions on weight and cost, play an important role that must be properly accounted for. In this
framework, some early choices could be revisited and subject to aerodynamic criteria checking, so
that an iterative process occurs until a satisfactory preliminary design is obtained. A second
preliminary step, distinct from the 1D procedure, is the two-dimensional design (2D), which include
both cascade and through flow models, from which a characterization of both design and off-design
multi-stage compressor performance can be carried out after some iterations, if necessary. In this
case, both direct and inverse design methodologies have been successfully applied.

Figure 10.5 Compressor Design Flow Chart

Numerical optimization strategies may be of great help in this case as the models involved are
relatively simple to run on a computer. Often an optimization involves coupling a prediction tool, e.g.
a blade to blade solver and/or a through flow code, and an optimization algorithm which assists the
designer to explore the search space with the aim of obtaining the desired objectives. Finally, a fully
three-dimensional (3D) design is carried out including all the details necessary to build the
aerodynamic parts of the compressor. In this phase, some design intervention is needed to account
for the real three-dimensional, viscous flows in the stages, especially tip clearance, secondary flows
and casing treatment for stall delay. This is usually carried out using CFD models, where the running
blade is modeled in its actual deformed shape, analyzed and, if necessary optimized. While
traditional 3D analyses are aimed at evaluating and improving compressor performance of a single
stage, the recent availability of powerful computers makes the analyses of multistage compressors
an affordable task for most industries. Most advanced CFD computations include evaluation of
complex unsteady effects due to successive full-span rotor-stator interaction315.

315 Benini, E., “Advances in Aerodynamic Design of Gas Turbines Compressors “, University of Padova Italy.
249

10.3.4 Preliminary Design Techniques (1D)


A simple mean or pitched line 1D method usually forms the basis of a preliminary design. For a given
design condition, the compressor total pressure ratio is known from which the total number of
compressor stages can be estimated (see Figure 10.6). With this respect, the designer can use
statistical indications based on typical values of admissible peripheral speeds and stage loading. This
is very useful for estimating the
preliminary stage pressure ratio once
the range for the other functional
parameters has been settled. Result of
stage stacking consists in the flow
path definition, from which the
distribution of stage parameters
along the mean radii can be obtained.
Because the stacking procedure is
intrinsically iterative, a loop is
required to satisfy all the design
objectives and constraints. As a first
check, the axial Mach distribution
along the stages must be calculated
and a value not exceeding 0.5 is
tolerated for both subsonic and
transonic stages. By imposing such a
constraint, the values of stage area
passage can be derived from the Figure 10.6 Preliminary Estimation of Number of Stages in
continuity equation316. Next, the Compressor
values of the hub-to-tip ratios must be
defined. To this end, it is worth recalling that such value comes from a trade-off between
aerodynamic, technological and economic constraints. For inlet stages, values between 0.45 and 0.66
can be assigned, while outlet stages often are given a higher value, say from 0.8 to 0.92, in order not
to increase the exit Mach number (a condition that is detrimental for pneumatic combustor losses).
Despite its relative simplicity, mean line 1D methods based on stage-stacking techniques still play an
important role in the design of compressor stages. Recent works includes a numerical methodology
used for optimizing a stator stagger setting in a multistage axial-flow compressor environment
(seven-stage aircraft compressor), based on a stage-by-stage model to 'stack' the stages together with
a dynamic surge prediction model. The absolute inlet and exit angles of the rotor are taken as design
variables. Analytical relations between the isentropic efficiency and the flow coefficient, the work
coefficient, the flow angles and the degree of reaction of the compressor stage were obtained.
Numerical examples were provided to illustrate the effects of various parameters on the optimal
performance of the compressor stage317.
10.3.5 Through Flow Design Techniques (2D)
Through flow design allows configuring the meridional contours of the compressor, as well as all
other stage properties in a more accurate way compared to 1D methods. They make use of cascade
correlations for total pressure loss/flow deviation and are based on through flow codes, which are
two-dimensional inviscid methods that solve for axisymmetric flow (radial equilibrium equations)
in the axial-radial meridional plane (Figure 10.7)318. A distributed blade force is imposed to
produce the desired flow turning, while blockage factor that accounts for the reduced area due to

316 Same as previous.


317 Same as previous.
318 Benini, E., “Advances in Aerodynamic Design of Gas Turbines Compressors “, University of Padova Italy.
250

blade thickness and distributed frictional force representing the entropy increase due to viscous
stresses and heat conduction can be incorporated. Three methods are basically used for this purpose:

• Streamline Curvature Methods SCM (Novak, 1967),


• Matrix Through Flow Methods (MTFM) (Marsh, 1968) and
• Streamline Through Flow Methods (STFM) (Von Backström & Rows, 1993).

Figure 10.7 Optimization Procedures Proposed in [Massardo et al.]

The SCM has the advantage of simulating individual streamlines, making it easier to be implemented
because properties are conserved along each streamline but is typically lower compared to the other
methods. On the other hand, MTFM uses a fixed geometrical grid, so that streamline conservation
properties cannot be applied. However, despite stream function values must be interpolated
throughout the grid, the MTFM is numerically more stable than SCM. Finally STFM is a hybrid
approach which combines advantages of accuracy of SCM with stability of MTFM. These methods
have recently been made more realistic by taking account of end-wall effects and span wise mixing
by four aerodynamic mechanisms: turbulent diffusion, turbulent convection by secondary flows,
span wise migration of airfoil boundary layer fluid and span wise convection of fluid in blade wakes.
As a result of the application of through flow codes, the compressor map in both design and off design
operation can be obtained exhibiting high accuracy319. Remarkable developments in the design
techniques have been obtained using such codes. Among others, [Massardo] described a technique
for the design optimization of an axial-flow compressor stage. The procedure allowed for
optimization of the complete radial distribution of the geometry, being the objective function
obtained using a through flow calculation (see Figure 10.7). Some examples were given of the
possibility to use the procedure both for redesign and the complete design of axial-flow compressor
stages.
10.3.6 Detailed Design Techniques (3D)
10.3.6.1 Direct Methods

319 Same Source – See Previous.


251

Advanced optimization techniques can be of great help in the design of 3D compressor blades when
direct methods are used320. These are usually very expensive procedures in terms of computational
cost such that they can be profitably used in the final stages of the design, when a good starting
solution, obtained using a combination of 1D and/or 2D methods, is already available. Moreover,
large computational resources are necessary to obtain results within reasonable industrial times.
Examples of 3D designs of both subsonic and transonic compressor blading’s are today numerous in
the open literature. For numerical optimization, searching direction was found by the steepest decent
and conjugate direction methods, and it was used to determine optimum moving distance along the
searching direction. The object of present optimization was to maximize efficiency. An optimum
stacking line was also found to design a custom-tailored 3D blade for maximum efficiency with the
other parameters fixed. The method combined a parametric geometry definition method, a powerful
blade-to-blade flow solver and an optimization technique (breeder genetic algorithm) with an
appropriate fitness function. Particular effort has been devoted to the design of the fitness function
for this application which includes non-dimensional terms related to the required performance at
design and off-design operating points. It has been found that essential aspects of the design (such as
the required flow turning, or mechanical constraints) should not be part of the fitness function, but
need to be treated as so-called "killer" criteria in the genetic algorithm. Finally, it has been found
worthwhile to examine the effect of the weighting factors of the fitness function to identify how these
affect the performance of the sections321.
A multi-objective design optimization method for 3D compressor rotor blades was developed by
[Benini, 2004], where the optimization problem was to maximize the isentropic efficiency of the rotor
and to maximize its pressure ratio at the design point, using a constraint on the mass flow rate. Direct
objective function calculation was performed iteratively using the 3D Navier-Stokes equations and a
multi-objective evolutionary algorithm featuring a special genetic diversity preserving method was
used for handling the optimization problem. In this work, blade geometry was parameterized using
three profiles along the span (hub, mid span and tip profiles), each of which was described by camber
and thickness distributions, both defined using Bezier polynomials. The blade surface was then
obtained by interpolating profile coordinates in the span direction using spline curves. By specifying
a proper value of the tangential coordinate of the first mid span and the tip profiles control point with
respect to the hub profile, the effect of blade lean was achieved. Results of tip profiles control point
with respect to the hub profile and the effect of blade lean was achieved. Performance enhancement
severe shock losses (Figure 10.7). Computational time was enormous, involving about 2000 CPU
hours on a 4-processor machine.
10.3.6.2 Inverse Methods
In the last two decades, 3D inverse design methods have emerged and been applied successfully for
a wide range of designs, involving both radial/mixed flow turbo-machinery blades and wings. Quite
a new approach to the 3D design of axial compressor blading has been recently proposed by [Tiow
2002]. In this work, an inverse method was presented which is based on the flow governed by the
Euler equations of motion and improved with viscous effects modeled using a body force model.
However, contrary to the methods cited above, the methodology is capable of providing designs
directly for a specific work rotor blading using the mass-averaged swirl velocity distribution.
Moreover, the methodology proposed by [Tiow], joins the capabilities of an inverse design with the
search potential of an optimization tool, in this case the simulated annealing algorithm. The entire
computation required minimal human intervention except during initial set-up where constraints
based on existing knowledge may be imposed to restrict the search for the optimal performance to a
specified domain of interest.

320 Same Source – See Previous.


321 Benini, E., “Advances in Aerodynamic Design of Gas Turbines Compressors “, University of Padova Italy.
252

Two generic transonic designs have been presented, one of which referred to compressor rotor,
where loss reductions in the region of 20 per cent have been achieved by imposing a proper target
surface Mach number which resulted in a modified blade shape. Figure 10.8 comparison of blade
loading distributions of an original supersonic blade, a new design (prescribed by inverse mode), and
a reference blade (R2-56 blade) for a given pressure ratio (left); comparison of passage Mach number
distributions at 95% span. Results
showed that an optimum combination
of pressure-loading tailoring with
surface objective can lead to a
minimization of the amount of sucked
flow required for a net performance
improvement at design and off-design
operations. By prescribing a desired
loading distribution over the blade the
placement of the passage shock in the
new design was about the same as the
original blade. However, the passage
shock was weakened in the tip region
where the relative Mach number is
high.
10.3.7 Concluding Remarks
Continuous effort is currently being
spent in building advanced design
techniques able to tackle the problem Figure 10.8 Comparison of Blade Loading Prescribed by
efficiently, cost-effectively and Inverse Mode
accurately. Plenty of design
optimization techniques has been and are being developed including standard trial and-error 1D
procedures up to the most sophisticated methods, such as direct or indirect methods driven by
advanced optimization algorithms and CFD. Advanced techniques can be used in all stages of the
design. In the field of 1D, or mean line methods, correlation-based prediction tools for loss and
deviation estimation can be calibrated and profitably used for the preliminary design of multistage
compressors. 2D methods supported by either through-flow or blade-to-blade codes in both a direct
and an indirect approach, can be used afterwards, thus leading to a more accurate definition of the
flow path of both meridional and cascade geometry. To enhance the potentialities of such methods,
optimization algorithms can be quite easily used to drive the search toward optimal compressor
configurations with a reasonable computational effort. Detailed 3D aerodynamic design remains
peculiar of single stage analyses, although several works have described computations of multistage
configurations, either in steady and unsteady operations. However, the latter is an approach suitable
for verification and analysis purposes, thus with a limited design applicability. The 3D design
optimization techniques can realistically be used if local refinement of a relatively good starting point
is searched for. On the other hand, if more general results are expected, simplified design methods
are mandatory, such as those based on supervised learning procedures, where surrogate models of
the objective functions are constructed. Other very promising techniques include adjoin methods,
where the number of design iteration can be potentially reduced by an order of magnitude if local
derivatives of physical quantities with respect to the decision variables are carefully computed322.

322 Benini, E.,” Three-Dimensional Multi-Objective Design Optimization of a Transonic Compressor Rotor”.
Journal of Propulsion and Power, Vol. 20, No. 3 (May/June), pp.559-565, ISSN: 0748-4658-2004.
253

10.4 Case Study 2 – Turbine Airfoil Optimization using Quasi 3D Analysis Codes
Turbine airfoil design has long been a domain of expert designers who use their knowledge and
experience along with analysis codes to make design decisions. The turbine aerodynamic design is a
three-step process that is pitch line analysis, through-flow analysis, and blade-to-blade analysis, as
depicted in Figure 10.9. In the pitch line analysis, flow equations are solved at the blade pitch, and
a free vortex assumption is used to get flow parameters at the hub and the tip. Using this analysis the
flow path of the turbine is optimized, and number of stages, work distribution across stages, stage
reaction, and number of airfoils in each blade row are determined. In the through-flow analysis, the
calculation is carried out on a series of meridional planes where the flow is assumed to be
axisymmetric and the boundary conditions of each stage are determined. The axisymmetric through-
flow method allows for variation in flow parameters in the radial direction without using the free
vortex assumption and accounts for interactions between multiple stages. In the blade-to-blade
analysis, airfoil profiles are designed on quasi-3D surfaces using a computational fluid dynamics
code.

Figure 10.9 Turbine Design Process

The design of airfoil profiles involves slicing the blade on quasi-3D surfaces, designing each section
separately, and stacking the sections together to obtain a sooth radial geometry. The objective of
airfoil design is to define the airfoil shape so as to ensure structural integrity and minimize losses.
The primary sources of losses in an airfoil are profile loss, shock loss, secondary flow loss, tip
clearance loss, and end-wall loss. Profile loss is associated with boundary layer growth over the blade
profile causing viscous and turbulent dissipation. This also includes loss due to boundary layer
separation because of conditions such as extreme angles of incidence and high inlet Mach number.
Shock losses arise due to viscous dissipation within the shock wave which results in increase in static
pressure and subsequent thickening of the boundary layer, which may lead to flow separation
downstream of the shock. End-wall loss is associated with boundary layer growth on the inner and
outer walls on the annulus. Secondary flow losses arise from flows, which are present when a wall
boundary layer is turned through an angle by an adjacent curved surface. Tip clearance loss is caused
by leakage flows in the tip clearance region of the rotor blade, where the leaked flow fails to
254

contribute to the work output and also interacts with the end-wall boundary layer. The objective of
the design is to create the most efficient airfoil by minimizing these losses. This often requires
trading-off one loss versus another such that the overall loss is minimized.
To compute all these losses a 3D viscous analysis is required; however, due to the computational load
of such a code, a quasi-3D analysis code is often used in the design process. Thus the impact of the
blade geometry on 3D losses cannot be determined and only 2D losses can be minimized, that is,
profile and shock losses. A viscous quasi-3D analysis though less computationally intense is still too
expensive for use in design optimization, and an inviscid quasi-3D code is used instead. Consequently,
viscous losses are not computed from the analysis code and airfoil performance is gauged by the
characteristics of the Mach number distribution on the blade surface. The most practical formulation
for low-speed turbine airfoil designs still remains the direct optimization formulation based on 2D
inviscid blade-to-blade solvers. This work automates the direct design process as described in the
next section323.
10.4.1 Parametric Representation of Airfoil Design Process
The parametric representations of the airfoils used in this work are based on the standard design
tools and practices. There are separate
models for the high-pressure and low
pressure turbine blades. The high-
pressure turbine blades are subject to
very high temperatures and need to be
cooled. The parametric representations
of the airfoils used in this work are based
on the standard design tools and
practices. There are separate models for
the high-pressure and low pressure
turbine blades. The high-pressure
turbine blades are subject to very high
temperatures and need to be cooled. As
a result, these airfoils are made thick to (A) High Pressure Airfoil
accommodate cooling passages inside
the blades. For such thick airfoils,
suction and pressure surfaces need to be
manipulated independently of each
other. So the airfoil is represented as a
combination of two separate curves, one
for the pressure side and the other for
the suction side (see Figure 10.10 (A))
Bezier curves are well suited for these
airfoils. Low- pressure airfoils, on the
other hand, have lower thermal stresses,
are much longer, and have a lower speed
of rotation compared to high-pressure
airfoils. These airfoils are usually very
thin and the two-surface model does not (B) Low Pressure Airfoil
work very well as it is very difficult to
vary the pressure and suction surfaces Figure 10.10 Parametric Representation of an
independently and still maintain a Airfoil

323
Goel, Sanjay,” Turbine Airfoil Optimization Using Quasi-3D Analysis Codes”, University at Albany, USA.
255

smooth thickness distribution. For such Geometry


airfoils, a mean line and thickness Definition
Variables
representation is used in which a thickness
Angle of line joining leading
distribution is super imposed on the mean line
Stagger & trailing edge of the airfoil to
of the airfoil as shown in Figure 10.10 (B). In
axial
this representation, the mean line and the
Maximum thickness of the
thickness distribution can be varied Tmaxx
airfoil
independently, and good control of the
thickness distribution is obtained. Here we C1 point of maximum thickness
discusses optimization of low-pressure turbine C2 trailing edge included angle
blades with the following parameters in Table
10.2. C3 curvature of mean line
10.4.2 Constraints and Problem curvature near leading edge
Ratu
Formulation on upper surface
Constraints are imposed on the airfoil curvature near leading edge
Ratl
geometry to ensure that the airfoil is on lower surface
manufacturable and structurally feasible as Pcttle incidence angle
well as for ensuring high aerodynamic
efficiency. The structural and manufacturing Ti leading edge bluntness
constraints are based on the airfoil geometry ellipse ratio of the
and the aerodynamic constraints are derived E approximate ellipse fitted in
from the Mach number distribution on the the nose
airfoil. There aerodynamic constraints are
defined, that is, peak-exit-ratio, peak-location, Table 10.2 Airfoil Geometry Parameters
and inlet-valley-ratio. These are listed below
and can be interpreted from the Mach number distribution shown in Figure 10.11. Peak-exit-ratio
is defined as the ratio of the peak Mach number on the suction surface to the Mach number at the
trailing edge of Constraints are imposed on the airfoil geometry to ensure that the airfoil is
manufacturable and structurally feasible as well as for ensuring high aerodynamic efficiency.
The structural and manufacturing constraints are based on the airfoil geometry and the aerodynamic
constraints are derived from the Mach number distribution on the airfoil. There aerodynamic
constraints are defined, that is, peak-exit-ratio, peak-location, and inlet-valley-ratio. These are listed
below and can be interpreted from
the Mach number distribution
shown in Figure 10.11. Peak-exit-
ratio is defined as the ratio of the
peak Mach number on the suction
surface to the Mach number at the
trailing edge of the blade. This is a
measure of flow acceleration on
the unguided portion of the airfoil
(between the throat and the
trailing edge). A very high turning
on the unguided portion of the
airfoil can lead to separation of
flow or the formation of a shock on
the trailing edge. By putting a
constraint on the maximum peak-
exit-ratio, chances of separation Figure 10.11 Sample Mach Number Distribution
256

are minimized. Peak-location is the normalized location of the peak Mach number on the suction side.
It is desirable to have an increasing Mach number as far along on the suction side as possible to
prevent a thickening of the boundary layer. Imposing a constraint on which allows the peak to occur
after 65% of the blade width guards against upstream diffusion and helps in achieving a smooth
accelerating Mach number on the suction side. Inlet-valley-ratio is the ratio of the Mach number at
the inlet of the airfoil on the pressure side, to the minimum Mach number on the pressure side. This
constraint controls the diffusion near the inlet on the pressure side and restricts the thickening of the
boundary layer, reducing chances of flow separation. Constraints are also imposed on:

• The curvature change on the unguided portion of the airfoil (unguided turning)
• The difference between the blades mean line angle and the flow angle at the trailing edge
(over turning), and
• The difference between the inlet angle and the metal angle at the inlet (Δ1).

These additional checks further ensure that the designed airfoil stays within design practice
guidelines324. To ensure mechanical and structural feasibility, constraints are imposed on the blade
geometry. The primary geometry parameters are cross section area, maximum thickness of airfoil,
wedge angle, and nose radius. In cooled airfoils, the constraints on the geometry stem from the
necessity to construct cooling channels in the airfoil; these constraints are dictated by manufacturing
requirements. In low-pressure airfoils,
Design Lower Upper Initial Final
these constraints are primarily driven by
s reassess and manufacturing limitations. Variables Bound Bound Value Value
Most of these constraints have soft limits C 1 0.2 0.5 0.35 0.35
on them; that is, it is best to have the C2 0.25 0.75 0.5 0.632
responses within a given range, beyond a C3 0.25 0.75 0.5 0.569
threshold of the range a penalty that Tmaxx 0.05 0.15 0.139 0.139
increases nonlinearly with increased
Stagger 8 40 31.643 39.464
violation of the constraint is added to the
objective function. The objective function Pcttle -0.25 0.25 0 0
includes the performance metrics and Ratl 0 4 1.25 2.703
constraints where the violations are Ratu 0 4 2.59442 2.727
included via penalty functions. These Ti 0 1 0.5 1
factors can vary for different problems E 1 5 3 2.044
based on the requirements of the specific
problem. The design variables and typical Table 10.3 Airfoil Design Variables
range of variations are listed
Table 10.2 in and the constraints imposed on the problem are listed in Table 10.3. During the
design of an airfoil, multiple sections are designed concurrently, and the objective function is a sum
of the objective functions of all the cross sections being designed. Constraints for all the sections are
also included in the problem formulation. Polynomial fits are used to represent the radial variation
of the design variables; thus the objective function becomes a combination of the coefficients of the
fits across multiple sections rather than individual parameters for each section. A second-order
polynomial fit is used in the formulation; so corresponding to each metric we have three coefficients.
The solution to the problem can be attempted using a variety of optimization techniques including
numerical optimization, genetic algorithms, simulated annealing, and heuristic search. In the current
investigation, the BFGS variable metric method implemented in an optimization code ADS was used.

324 Goel, Sanjay,” Turbine Airfoil Optimization Using Quasi-3D Analysis Codes”, University at Albany, USA.
257

A one-dimensional search technique was used in which the search was bounded followed by use of
polynomial interpolation.
10.4.3 Quasi-3D CFD Analysis and Results
A quasi 3D CFD solver is used in the current investigation to analyze the flow on the airfoil, which is
an isentropic that uses the streamline curvature method that computes the Mach Number/Pressure
distribution on the airfoil surface 325. In the absence of a viscous code, designers usually estimate the
quality of the airfoil by visually examining the Mach number distribution obtained from an in-viscid
quasi 3D CFD solution. Since optimization techniques are driven by a numerical value of the objective
function, and the visual perspective of the designer is the only proven metric available, it must be
captured in a suitable numerical algorithm to provide a measure of quality of an airfoil. The current
work employs curve fitting coupled with design heuristics to compute quality metrics from the Mach
number distribution and the airfoil geometry. These metrics are weighted for different designs based
on individual designer preferences. Primary evaluation metrics that have been defined are diffusion,
deviation, incidence deviation, and leading edge crossover. A physical interpretation of these metrics
is presented below. Diffusion is defined as the deceleration of the flow along the blade surface. It is
measured as the cumulative aggregate of all flow diffusions at each point along the airfoil surface. As
the flow diffuses, the boundary layer thickens, and the momentum loss in the boundary layer
increases. In this case, the increased drag causes a significant loss of momentum; flow separation
may result, causing much larger losses. Thus, the objective of the design is to minimize the diffusion
effect. Since the impact of diffusion on the pressure and suction sides is different, separate terms are
defined for the suction and pressure sides. In the test case presented here, a low-pressure turbine
nozzle is optimized. The flow-path of the low-pressure turbine used in the investigation is shown in
Figure 10.12. The radial distances in the figure are measured with reference to the centerline of
the engine and the axial distances are measured with reference to a point upstream of the first stage
of the turbine. The horizontal lines in the figure represent the streamlines of the flow. Thirteen
streamlines are shown, the top and bottom of which coincide with the casing and the hub
respectively. The vertical lines represent the edges of the blade rows and the location of the frame.
The turbine has six stages, each stage composed of two blade rows. The first blade row consists of

Figure 10.12 Flow path of the turbine

325 Goel, Sanjay,” Turbine Airfoil Optimization Using Quasi-3D Analysis Codes”, University at Albany, USA.
258

nozzles and the second blade row consists of buckets. The stages are numbered from 1 to 6 in the
Figure 5.5.
In the current investigation, stage 5 nozzle was designed using sections from five streamlines equally
spaced along the blade span (hub to
tip). Figure 10.13 Shows the
approximate locations of the
streamlines for an airfoil in which the
first and the last streamlines are shown
at the hub and tip. In reality however
streamlines at 5% and 95% span were
used instead of streamlines directly on
the hub and tip because Mach number
distributions very close to the end walls
are distorted by the end wall effects and
not representative of the flow away
from the walls. The starting solution for
the test case was obtained by
estimating the airfoil shape based on Figure 10.13 Schematics of an airfoil showing
shapes of similar airfoils designed in stream lines along the radial direction
the past. All the Mach number and
airfoil geometry plots use the same reference radial and axial locations as shown in Figure 10.14.
To ensure slope and curvature smoothness of the geometry, second- order polynomials were used to
represent the radial distribution of geometry parameters.
Thus there are three design variables for each geometry
parameter, that is, C0, C1, and C2. These are the coefficients of the
2nd polynomial representing the geometry parameter. The
efficient of the fit match well with the starting design since the
design is based on a previously designed airfoil. Subsequently the
smoothness is maintained since the parameters are not changed
directly but rather the coefficients of the polynomials are varied.
The geometry parameters which describe the low-pressure
turbine airfoil geometry are Stagger, Tmaxx, C1, C2, C3, Ratu, Ratl,
Pcttle, ti, and E. These geometry parameters are varied within
limits typically prescribed in design practice and on the basis of
prior experience and manufacturing limitations. The limits for
these parameters are described along with the results for each Figure 10.14 3D model of
specific test case. an airfoil showing the
passage between adjacent
10.4.4 Concluding Remarks
airfoils
Here we presented a mathematical formulation for design of
turbine airfoils using 2D geometry models and 2D inviscid
analysis codes. The reduced computational complexity of the new formulation compared to 3D
viscous analysis makes the airfoil design problem amenable to the use of formal optimization
methods. The paper presents results from design of a low-pressure turbine nozzle. There are three
primary contributions of this work:

• A numerical metric for emulating designer judgment in evaluation of airfoil Mach number
Distribution.
• An optimization formulation for design of airfoil sections.
259

• A methodology which allows design of 3D airfoils by simultaneous design of multiple 2D


sections.

Designer heuristics are computed using curve fits and error norms. A set of penalty functions has
been defined which allows for flexible constraint boundaries and influence constrained variables
even within constraint limits. In the new approach multiple two-dimensional sections of the airfoil
are designed with constraints on radial smoothness using polynomial fits on the parametric
geometry variables in the radial direction (Figure 10.14)326. Airfoil design is a labor intensive,
repetitive, and cumbersome task for the designers and is a bottleneck for both the design cycle and
rapid generation of inputs for complex multistage analyses. Automating the design process
significantly cuts down the design cycle time and facilitates the task of running multistage analysis
by rapidly generating airfoil geometries. While designing an airfoil, it is hard to establish the
existence of a unique optimum. Multiple evaluation criteria which are weighted together to define
the objective function and the relative importance of these are determined based on designer
experience. Furthermore, the analysis codes are not exact, and even with precisely defined quality
metrics, a significant margin of error remains. In manual design the evaluation criteria are implicitly
considered by the designer, with weighting factors based on past experience and individual biases.
Subjectivity is introduced into the design process since the evaluation criteria for the design are
partially based on heuristics abstracted from designer experiences. Thus in order to completely
understand the results of airfoil optimization, an evaluation of the qualitative changes to the design
is essential after the optimization is completed. Over time as the metrics to evaluate airfoil design
become more acceptable, a standard metric will emerge, till such time designers will need to tinker
with the weights to suit their own preferences 327

10.5 Case Study 3 – 2D Design Optimization of Turbine Blade in Quasi-Periodic


Unsteady Flow Problems Using a Harmonic Balance Method
The aim here is to assess the capability of the Harmonic Balanced based design method for fully-
turbulent flows [Rubino et al.]328 for turbomachinery application. With the aid of the experimental
setup329, the unsteadiness in the wake of an upstream blade row is approximated by a moving bar, as
depicted in Figure 10.15. The moving bars are located at xb/l = 0.7 upstream of the cascade inlet
plane, having a velocity vb=21.4m/s parallel to the inlet. A schematic representation is shown in
Figure 10.15, and the main operating conditions for the test case are reported in [Rubino et al.].
The two-dimensional flow domain is discretized with approximately 40000 elements and the Roe
scheme is selected for the convective flux discretization. A suitable spacing of quadrilateral elements
is used to cluster the near wall cells such that y+ is less than 1. This test case is a benchmark for the
study of laminar-to-turbulent transition. However, since the present work aims to assess the
methodology for design optimization only, the computations are performed assuming fully-turbulent

326 Goel, Sanjay,” Turbine Airfoil Optimization Using Quasi-3D Analysis Codes”, University at Albany, USA.
327 Goel, Sanjay,” Turbine Airfoil Optimization Using Quasi-3D Analysis Codes”, University at Albany, USA.
328 A. Rubinoa, M. Pini, P.Colonna, T. Albring, S. Nimmagadda, T. Economon, J. Alonso, “ Adjoint-based fluid

dynamic design optimization in quasi-periodic unsteady flow problems using a harmonic balance method”,
https://doi.org/10.1016/j.jcp.2018.06.023.
329 P. Stadtmüller, L. Fottner, “A test case for the numerical investigation of wake passing effects on a highly

loaded LP turbine cascade blade”, ASME Turbo Expo 2001, Power for Land, Sea, and Air, ASME, 2001.
260

conditions, employing the SST turbulence model330. In order to calculate the cascade performance,
the total pressure loss coefficient is evaluated a

〈Ptot,1 〉∂Ω1 − 〈Ptot,2 〉∂Ω2


ξP =
〈Ptot,2 〉∂Ω2 − 〈P2 〉∂Ω2
Eq. 10.2
where <Ptot,1>∂Ω1 and <Ptot,2>∂Ω2 are the
inlet and outlet total pressure averaged
over their corresponding boundary,
respectively. <P2>∂Ω2 is the average static
pressure at the cascade outlet. The
averages at the boundaries are calculated
using a mixed-out averaging
procedure331. First, a validation is
performed using the experimental data of
the turbine cascade operating at steady
state. The simulation results show a very
good agreement with the experimental
data. As expected, the main deviation
between CFD and experiments occurs at
about x/l =0.75, possibly due to transition
not being accurately modeled. For this Figure 10.15 Schematic Geometry of theT106D-IZ
test case, two configurations are Turbine Cascade
considered for design:
1. Spatially non-uniform, time-dependent inlet boundary condition;
2. Spatially uniform, time-dependent inlet boundary condition.
The terminology OptC1 and OptC2 is used to refer to the first and the second shape optimization
problem, respectively. Here, we investigate the Non-Uniform case (optC1). Readers are encourage to
consult the [Rubino et al.]332 for details on the second case.
10.5.1 OptC1 Configuration
In this configuration, an inlet boundary condition is imposed in order to reproduce the wakes
generated by the moving bars. The imposed values of the total pressure, temperature, and flow
direction at the boundary are interpolated from the results of a steady state simulation of the flow
past the bars. With this boundary condition, only multiples of the blade passing frequency are
expected. The fundamental blade passing frequency, given a ratio between the blade pitch and the
bar pitch yp/yb = 3, is defined as
3𝜈𝑏
f1 =
𝑦𝑏
Eq. 10.3

330 F. Menter, M. Kuntz, R. Langtry, “Ten years of industrial experience with the SST turbulence model, in:
Turbulence, Heat and Mass Transfer”, vol. 4 (1), 2003, pp.625–632.
331 A. Prasad, “Calculation of the mixed-out state in turbomachine flows”, J. Turbomachinery. 127(3) (2004).
332 A. Rubinoa, M. Pini, P.Colonna, T. Albring, S. Nimmagadda, T. Economon, J. Alonso, “ Adjoint-based fluid

dynamic design optimization in quasi-periodic unsteady flow problems using a harmonic balance method”,
https://doi.org/10.1016/j.jcp.2018.06.023.
261

To verify the HB solution, a second-order time-accurate URANS simulation using the dual time
stepping method is per-formed with a time-step 150xsmaller than the lowest period (1/f1). The total
pressure loss coefficient from this simulation is compared and HB solution obtained with 3, 5, and 7
time instances. The selected time instances correspond to the solution for the frequency vectors ωN3
= [0, ± ω0], ωN5 = [0, ± ω0, ± 2ω0] and ωN7 = [0, ± ω0 , ± 2ω0, ± 3ω0]. The resolved frequencies are,
therefore, multiples of the fundamental blade passing frequency only. The total pressure loss
coefficient, defined in Eq. 10.2, as function of time, and is obtained by spectral interpolation of the
harmonic balance result. The RMSE of the total pressure loss coefficient for the solution obtained
with 5 time instances is equal to 0.010. The harmonic balance solution obtained with 5 time instances
is about 9x faster than the time-accurate solution calculated over a total simulation time of five
periods, which includes the initial transient before reaching convergence to a periodic flow field
solution. 5 time instances are used for shape optimization, as a trade-off between accuracy and
computational cost. In Figure 10.18 - a , b, and c, the Mach number contours from the HB simulation
are reported for 3 different time instances with the simulation period given by T = 1/f 1. The results
show the bar wakes entering the cascade and a separation area occurring at about x/l = 0.7.
10.5.2 Optimization Problem and Results
The shape optimization problem of the cascade configuration is considered. It can be expressed as


Minimize ζP (𝐔𝐧 , 𝐗 𝐧 , 𝛂)
α
Subject to α out < α out,0 + 4 and δt = δt0
𝐔n < Gn n = 1 , 2, , , , , , N and 𝐗 𝑛 = M𝑛
Eq. 10.4
where the time-averaged total pressure loss coefficient ζP, obtained from Eq. 10.2, is selected as
objective function. Inequality constraints on the absolute exit flow angle (α out) and trailing edge
thickness (δt) are imposed. The optimization is performed using an ensemble of 16 geometrical

Figure 10.16 Shape Optimization History of the Total Pressure Loss Coefficient and Comparison
Between Baseline and Optimized Blade Profile (OptC1)
262

design parameters αbased on a


free-form deformation (FFD)
approach333. The gradients of the
objective function are again
obtained with the proposed
adjoint technique and compared
with the same gradients obtained
by second-order central finite
differences (FD). The results of
this comparison as reported in
[Rubino et al.]334 showing
excellent agreement between AD
and FD gradients (RMSE = 2
x10−5). The ratio between the
computational time of the adjoint
solution and the primal flow Figure 10.17 Total Pressure Loss Coefficient Evolution in time
Calculated with URANS Simulation for both the Baseline and the
solution is approximately 1.7.
Optimized Configuration
Figure 10.17 shows that the

Base line (a) Base line (b) Base line (c)

Optimized (d) Optimized (e) Optimized (f)

Figure 10.18 Mach Number Contours calculated at Three Different Time Instances with the HB
Method, Based on the OptC1 test case, for both the Baseline (a), (b), (c) and the Optimized (d), (e), (f)
Blade Profile

333 J. Samareh, Aerodynamic shape optimization based on free-form deformation, in: 10th AIAA/ISSMO
Multidisciplinary Analysis and Optimization Conference, 2004.
334 A. Rubinoa, M. Pini, P.Colonna, T. Albring, S. Nimmagadda, T. Economon, J. Alonso, “ Adjoint-based fluid

dynamic design optimization in quasi-periodic unsteady flow problems using a harmonic balance method”,
https://doi.org/10.1016/j.jcp.2018.06.023.
263

convergence of the optimization to the minimum objective is nearly reached after only 7 evaluations,
although satisfying the constraint requires more evaluations. Figure 10.17-b highlights that the
performance of the optimized blade is significantly improved, as the total pressure loss coefficient is
approximately 38%lower, while the constraint on the absolute outlet flow angle is satisfied. The
separation area, as seen in Figure 10.18, is considerably smaller with the optimized blade shape.
The unsteady optimization leads to a decrease in the peak of the total pressure loss coefficient of 44%
and a reduction of 54% of the signal amplitude in Figure 10.17. Furthermore, the objective function
spectrum obtained from a URANS simulation of the optimized blade (Figure 10.17) does not contain
additional frequencies when compared with the baseline configuration.

10.6 Case Study 4 - Using Shock Control Bumps To Improve Transonic Compressor
Blade Performance335
Citation : John, A., Qin, N., and Shahpar, S. (March 2, 2019). "Using Shock Control Bumps to Improve
Transonic Fan/Compressor Blade Performance." ASME. J. Turbomach. August 2019; 141(8):
081003. https://doi.org/10.1115/1.4042891
Shock control bumps can help to delay and weaken shocks, reducing loss generation and shock-
induced separation and delaying stall inception for transonic turbomachinery components, as
described by [John et al.]336. The use of shock control bumps on turbomachinery blades is
investigated here for the first time using 3D analysis. The aerodynamic optimization of a modern
research fan blade and a highly loaded compressor blade are carried out using shock control bumps

Figure 10.19 Tip Speed With Pronounced Negative Camber. From Prince – Courtesy of [Prince]

335 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic
Fan/Compressor Blade Performance”, GT2018-77065.
336 See Previous.
264

to improve their performance. Both the efficiency and stall margin of transonic fan and compressor
blades may be increased through the addition of shock control bumps to the geometry. It is shown
how shock induced separation can be delayed and reduced for both cases. A significant efficiency
improvement is shown for the compressor blade across its characteristic, and the stall margin of the
fan blade is increased by designing bumps that reduce shock-induced separation near to stall. Adjoint
surface sensitivities are used to highlight the critical regions of the blade geometries, and it is shown
how adding bumps in these regions improves blade performance. Finally, the performance of the
optimized geometries at conditions away from where they are designed is analyzed in detail.
10.6.1 Introduction and Motivation
Shocks are a major source of loss for transonic fans and compressors. They cause entropy generation,
boundary layer thickening and shock induced separation. The impingement of the shock on the blade
suction surface (and the resulting, strong, adverse pressure gradient) can cause the boundary layer
to detach, leading to larger blade wakes, reduced efficiency, lower blade stability and reduced stall
margin. Any method that can be used to alleviate shock strength (and the associated negative effects)
therefore has the potential to significantly improve transonic fan/compressor performance.
10.6.2 Shock Control for Turbomachinery & Literature Survey
Relatively little work on designing geometries directly to weaken the shock waves in transonic
turbomachinery components can be found in the literature, though it has been known for some time
that reducing the pre-shock Mach number of transonic compressors can improve their efficiency337.
It was clear to transonic compressor designers in the 70s and 80s that shock strength was increased
by the amount of convex curvature on the suction side between the leading edge and the shock338.
Nearly flat suction surfaces that minimized the expansion were therefore favored, with the next step
to try designs with concave curvature (often referred to as negative camber). Geometries with
negative camber result in gradual compression along the suction surface which weakens the shock.
The concave curvature of the blade surface and the reduction of flow area in the flow direction leads
to a deceleration of the supersonic flow through compression waves, and therefore a weaker shock.
[Prince]339 designed a rotor with pronounced negative camber. This lead to a rise in static pressure
along the suction surface prior to the shock as intended, but the resulting efficiency was
disappointing due to the strong shock on the pressure surface. [Ginder and Calvert]340 had more
success in designing a rotor with negative camber. With negative camber, the Mach number ahead of
the shock was reduced to 1.4 (compared to 1.5 for the traditionally designed blade) which drastically
reduced the amount of boundary layer separation and loss.
Recently, it was demonstrated by [John et al.]341 how the freeform shaping of a compressor blade can
improve blade efficiency by delaying and weakening the shock and reducing separation. The flexible
parameterization method used allowed an s-shaped design to be generated that included a pre-
compression geometry around mid-span. This s-shaped, pre-compression geometry is similar to the
negative camber designs described above. The effect of the pre-compression geometry on the shock
and separation is described in Figure 10.20. The current work proposes the use of shock control
bumps as an alternative method to reduce shock related loss to those described above. Shock control
bumps have the benefit that relatively small modifications to the original geometry are required to

337 Ginder, R., and Calvert, W., 1987. “The design of an advanced civil fan rotor”. Journal of turbomachinery,
109(3), pp. 340–345.
338 Cumpsty, N. A., 1989. Compressor aerodynamics. Longman Scientific & Technical.
339 Prince, D. C., 1980. “Three-dimensional shock structures for transonic/supersonic compressor rotors”.

Journal of Aircraft, 17(1), pp. 28–37.


340 Ginder, R., and Calvert, W., 1987. “The design of an advanced civil fan rotor”. Journal of turbomachinery,

109(3), pp. 340–345.


341 John, A., Shahpar, S., and Qin, N., 2017. “Novel compressor blade shaping through a free-form method”.

Journal of Turbomachinery, 139(8), p. 081002.


265

achieve the desired effect.

Figure 10.20 Schematic Of Shock Structures (A) Datum, (B) S-Shaped Design. Courtesy of [John et
al.]

10.6.3 Shock Control Bumps


Shock control bumps are bumps added to aerodynamic surfaces to alter the behavior of the shock
and improve aerodynamic performance. One of the earliest examples of 2D shock control bump usage
is in the design of the dromedary foil in the 1970s. This was a modified supercritical airfoil with a
bump added in an attempt to increase its drag-divergence Mach number. The ’hump’ was shown to
weaken the shock wave when implemented in the right position, acting as a localized precompression
ramp. This also demonstrated the importance of shock control bump positioning, as, if the bump was
misplaced, an increase in wave drag was seen. [Ashill et al.]342 found for a 2D airfoil a significant
reduction in drag could be achieved via the correct application of a shock control bump, however
when the shock position changed severe drag penalties were incurred due to secondary shocks and
separation being produced. [Drela and Giles]343 carried out numerical studies into shock control in
1987, describing the behavior of shock-induced separation. [Sommerer et al.]344 optimized shock
control bumps at various Mach numbers. They concluded that the bump height, width and position
of the bump peak are the key parameters. [Collins et al.]345 tested shock control bumps in a wind
tunnel, and analyzed the performance of shock control bumps at off-design conditions.
The EUROSHOCK II project began in 1996 and concluded that shock control bumps had the most
potential out of a range of shock control devices tested. A large amount of research was carried out
into shock control bumps, with both 2D and 3D analysis, although no optimization was undertaken.

342 Ashill, P., and Fulker, J., 1992. “92-01-022 a novel technique for controlling shock strength of laminar-flow
aerofoil sections”. DGLR BERICHT, pp. 175–175.
343 Drela, M., and Giles, M. B., 1987. “Viscous-inviscid analysis of transonic and low Reynolds number airfoils”.

AIAA journal, 25(10), pp. 1347–1355.


344 Sommerer, A., Lutz, T., and Wagner, S., 2000. “Design of adaptive transonic airfoils by means of numerical

optimisation”. In Proceedings of ECCOMAS, 2000, Barcelona.


345 Colliss, S. P., Babinsky, H., N¨ubler, K., and Lutz, T., 2016. “Vortical structures on three-dimensional shock

control bumps”. AIAA Journal, pp. 2338–2350.


266

[Qinet al.]346 first proposed 3D shock control bumps with a finite width, allowing additional design
complexity. They showed that 3D bump configurations were more robust than 2D bump designs
(where a 2D bump is extended continuously along the span).
The only use of a shock control bump on turbomachinery blades found in the literature is by
[Mazaheri and Khatibirad]347, who tested a 2D shock control bump on a (mid-span) section of the
NASA rotor 67 geometry. They added a bump modelled using the Hicks-Henne function. It was shown
how the interaction of the bump with the original wave structure resulted in a more desirable
pressure gradient, with a weaker compression wave fan and a more isentropic compression field.
The bump design was optimized and was shown to reduce the separation area at an off-design
condition. They describe how this may have the potential to improve the stall properties of the blade
section. Two optimizations were carried out, one at the design condition and another at 4% higher
rotational speed. Optimal bumps were produced for each condition, with an increase in efficiency of
0.67% for the on-design case and 2.9% in the off design case reported. The optimized geometry for
the design condition is shown in Figure 10.21.
The work by [Mazaheri & Khatibirad]348 demonstrated the benefit that bumps may provide at both
on and off-design conditions, and their potential to improve stall margin. The simplified 2D analysis
lacks accuracy
however as the
complex
behavior of
radial and
separated flow
cannot be
predicted. For a
thorough
understanding
of the potential
for the use of
shock control
bumps, 3D
analysis and the
design of 3D
Figure 10.21 Datum Geometry and Optimized Shock Control Bumps on The Mid-
bumps is
Section of Nasa Rotor 67- From Mazaheri et al..
needed to truly
assess their effect.
10.6.4 Test Case - NASA Rotor 37
The case studied here is NASA Rotor 37. This has a very strong shock wave (with a relative tip Mach
number of nearly 1.5) which causes large separation, decreasing the blade efficiency. It is a well-
documented case, having been extensively tested and simulated as part of a turbomachinery
validation study. It is a transonic rotor with inlet hub-to-tip ratio 0.7, blade aspect ratio 1.19, rotor
tip relative inlet Mach number 1.48 and rotor tip solidity 1.29. It has historically been a challenge for
CFD simulation. The very high pressure ratio, strong shock wave-boundary layer interaction, large
tip leakage vortex and highly separated flow mean that it poses challenges for turbomachinery

346 Qin, N., Wong, W., and Le Moigne, A., 2008. “Three dimensional contour bumps for transonic wing drag
reduction”. Proceedings of the Institution of Mechanical Engineers, Part G: Journal of Aerospace Engineering,
222(5), pp. 619–629.
347 Mazaheri, K., and Khatibirad, S., 2017. “Using a shock control bump to improve the performance of an axial

compressor blade section”. Shock Waves, 27(2), pp. 299–312.


348 See Previous.
267

solvers. Rotor 37 has been the subject of review articles that highlight the complexity of matching
experimental and computational measurements and the associated uncertainties. The CFD setup is
shown in Figure 10.22. At the inlet, a radial distribution of total pressure and temperature (based
on the original experimental values) is specified. The inlet turbulence intensity is 1%. At the outlet, a
value for circumferentially mixed-out and radially mean-mass capacity (non-dimensional mass flow)
is used. Periodic boundaries are used to represent full annulus flow. Stationary walls are treated as
adiabatic viscous walls and the rotational speed of the non-stationary portions of the domain is

Figure 10.22 The R37 CFD Domain Used – Courtesy of [John et al.]

1800.01rads􀀀1, as specified in the experiment. Rolls-Royce CFD solver Hydra is used for all of the
simulations presented here, using the Spalart-Allmaras turbulence model (fully turbulent). The 4.27
M cell mesh is generated by PADRAM, has y+ of the order of one on all surfaces with 30 cells in the tip
gap.
10.6.5 Validation
As previously alluded to, many studies have struggled when matching simulations of Rotor 37 to the
experiment. A wide range of work has been undertaken to investigate the discrepancy found between
simulation and experiment, with the primary work being the 1994 ASME/IGTI blind test case study
in which a range of codes were used to simulate the rotor, with no knowledge of the experimental
values. A large variation was seen between the different predictions, prompting analysis by
268

[Denton]349. Recent work has also been carried out by [Chima]350 and [Hah]351. The differences are
usually attributed to uncertainty in the experimental measurements, the lack of real geometry in the
simulations (e.g. the upstream hub cavity is usually missing) and also the difficulty in fully resolving
the complex flows. The pressure ratio agreement is reasonable across the characteristic, but the
efficiency prediction is about 2% below the experimental value at the design point (98% choke). This
matches the trend of previous results, where the better the PR prediction, the worse the efficiency
match. This ’trade-off’ has been seen in a range of previous simulations. Figure 10.23 gives the
radial profiles of total PR and efficiency at 98% of simulated choke compared to the experimental
values at 98% experimental choke. The radial trends have been captured fairly well, although there
is an offset from the experiment for both. The choke mass flow found in the simulations was
20.91kg/s, matching quite closely the experimental of 20.93kg/s.

Figure 10.23 Radial Profiles Vs Experimental Data – Courtesy of [ John et al.]

10.6.6 Flow Field For the Datum Case


Figure 10.24 shows the flow features of the datum NASA Rotor37 at design point. It can be seen
how the strong shock of Rotor 37 causes complex shock-boundary layer interaction and a large
shock-induced separation (this can be seen by the thickening of the boundary layer and wake shown
in Figure 10.24-b and the orange contour of zero axial velocity in Figure 10.24-a. At the point
where the shock impinges on the suction surface, its interaction with the boundary layer causes it to

349 Denton, J., 1997. “Lessons from rotor 37”. Journal of Thermal Science, 6(1), pp. 1–13.
350 Chima, R., 2009. “Swift code assessment for two similar transonic compressors”. In 47th AIAA Aerospace
Sciences Meeting including The New Horizons Forum and Aerospace Exposition, p. 1058.
351 Hah, C., 2009. “Large eddy simulation of transonic flow field in NASA rotor 37”. 47th AIAA Aerospace

Sciences Meeting including The New Horizons Forum and Aerospace Exposition, p. 1061.
269

separate and a large wake forms. It is at this design point that Rotor 37 will be optimized, as a
reduction in this separation could significantly increase blade efficiency.
10.6.7 Validation
Due to experimental data for this geometry not being available, simulation validation was carried out
using a similar fan blade geometry that has experimental data available. The related blade has very
similar performance parameters, and the simulation set up is identical. The results are given here. A
comparison of the simulations of this related blade against experimental data can be seen in 352. Both
the pressure ratio and efficiency curves match the experimental data well, though there is a slight
offset to the overall values and stall margin. The radial curves show good comparison to experimental
data, although the radial variation in efficiency is under predicted compared to the experiment.
Overall, the simulation compares well, lying within 1% across the range of flow rates.

Figure 10.24 (a) 3d Separation (Orange) On The R37 Geometry (Flow Right To Left), (b) Rel. Mach
No. Contour At 60% Span – Courtesy of [John et al.]

10.6.8 Blade Flow Features


To understand the behavior of this blade design and select a point at which to optimize the geometry,
the flow behavior for a range of flow rates was studied. Figure 10.25 shows the flow features of the
blade design as the flow rate is varied (see [John et al.]353. Point A is stalled. For proprietary reasons
the whole RR-FAN blade geometry cannot be shown, hence, flow behavior in just the region of
interest is shown in the following figures. The shock position on the blade surface moves towards the
LE as the operating point moves to the left on the characteristic. As the pressure ratio increases and
flow rate becomes lower, the strength of the shock increases and separation is caused towards stall.
It is this separation (highlighted in orange) that contributes to the full stall of the blade. It can be seen
that the shock-induced separation increases in magnitude and radial extent as the flow rate is
lowered until full separation eventually occurs. These near-stall operating points are a promising

352 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic
Fan/Compressor Blade Performance”, GT2018-77065.
353 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic

Fan/Compressor Blade Performance”, GT2018-77065.


270

area to investigate the benefit of shock control bumps. It is the shock-induced separation that is
responsible for limiting the operating range of the blade, and if this separation can be reduced then
it is expected that this will extend the stable working range of this fan.

Figure 10.25 Shock Region Flow Features For RR-Fan At Points A) A, B) B, C) C, D) D, E) E, F) F. Flow
Direction – Courtesy of [John et al.]

10.6.9 Adjoint Sensitivity Analysis


Adjoint sensitivity analysis is a useful tool that can be used to provide information on the sensitivity
of an objective function to changes in the geometry. Here, the adjoint sensitivity used is the
sensitivity of efficiency (as a percentage) to surface deformation (in mm) normal to the surface. This
can be used to inform which regions of the blade will have the greatest impact when modified, and
are therefore most important to control during an optimization. Hydra Adjoint354 is used to provide
the blade surface sensitivities: A primal Hydra simulation is first used to provide the flow solution,
followed by Hydra adjoint which calculates the flow-adjoint sensitivity and provides the sensitivity
of the objective function to changes in the flow. Once these two relatively expensive simulations are
completed, the mesh sensitivities are then mapped onto the surface. This finds the relationship
between changes in the flow to changes in the blade surface mesh. Combining these provides the
sensitivity (gradient) of the objective function (efficiency) to perturbations of the blade surface. The
adjoint surface sensitivity analysis for Rotor 37 at design point and RR-FAN at point D. It can be seen
that the most sensitive regions of both geometries are focused around the shock on the suction
surface. This indicates that geometry changes in this region will have a significant impact on the blade

354 Duta, M. C., Shahpar, S., and Giles, M. B., 2007. “Turbomachinery design optimization using automatic
differentiated adjoint code”. ASME Turbo Expo 2007: Power for Land, Sea, and Air, American Society of
Mechanical Engineers, pp. 1435–1444.
271

efficiency, and therefore if shock control bumps are applied here some benefit should be found. For
complete details, please consult the [John et al.]355
10.6.10 Shock Bump Parameterization & Optimization
The CST (Class Shape Transformation) method is used in this work to define the bump geometries.
The CST method uses Bernstein polynomials to create smooth (second derivative continuous)
contour bumps. For this project 3rd order CST bumps are used, constructed from four Bernstein
polynomials.
Controlling the
weighting (amplitude)
of these polynomials
modifies the bump
height and asymmetry.
Figure 10.26 shows
how the bump
geometry (solid black
line) to be added to the
blade surface is a sum of
the four Bernstein
polynomials (colored
dashed lines). The CST
bump parameterization
provides a high degree
of flexibility, enabling
the generation of Figure 10.26 Example 2d CST Bump (Solid Line) And The Four
Polynomials Used To Construct It (Dashed Lines) – Courtesy of [John et al.]
smooth, asymmetric
bumps in 2D and 3D.
The CST bump parameterization
technique was implemented inside of
the PADRAM geometry and meshing
software. The technique modifies each
2D radial section of the blade geometry,
adding a bump. The properties of these
2D bumps are smoothly interpolated in
the radial direction from control
sections. The resulting geometry is
controlled by the bump start and end
positions, the four Bernstein
polynomial amplitudes and the span-
wise distribution. This allows 3D
variation of the bumps in the radial
direction. Both continuous (where Figure 10.27 a) Example Individual Bump Geometry
bump amplitudes are smoothly And B) Example Continuous Bump Geometry – Courtesy of
interpolated radially) and individual [John et al.]
(where the bump amplitude returns to
zero periodically in the radial direction) CST bumps were tested. Examples of the blade with
individual and continuous bumps added is shown in Figure 10.27.

Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic
355

Fan/Compressor Blade Performance”, GT2018-77065.


272

During this work, a study was carried out (not detailed here for brevity) to compare the benefit of
using individual bumps (where a series of discrete bumps is added to the datum geometry in the
radial direction) with a continuous bump (note continuous bumps are still ’3D’ and their shape,
position and amplitude can vary in the radial direction). It was concluded that, for these cases, the
individual bumps needed to have greater amplitude than the continuous bumps to offer the same
benefit, leading to increased separation downstream of the bump position. The continuous bumps
tested offered greater benefit, and therefore only results using the ’continuous’ bump geometry
approach are presented here.
10.6.11 Optimization Method
In this work the Multi-point
Approximation Method (MAM) is
used for the optimization studies. It is
a gradient based method that uses
localized Design of Experiments (DoE)
and trust regions to efficiently search
through the design space. When using
MAM, an initial generation of
simulations (chosen by DoE) is carried
out around the start point. A response
surface is constructed for this region
and the sub-optimal point found. The
search is then moved to this point,
where a new generation is Figure 10.28 Spanwise Slice Of The Datum And Optimized
constructed and the process repeated R37 Geometries At 60% Span – Courtesy of [John et al.]
until the search converges on the
optimal design. The MAM method has
been shown to be an efficient and consistent approach for a wide range of highly constrained
optimization problems, working successfully for design
spaces made up of hundreds of parameters.
10.6.12 Rotor 37 Bump Optimization
For the Rotor 37 optimization, the bump geometry was
controlled at 5 radial heights (to allow radial variation of
the parameters) with the geometry smoothly interpolated
between the control stations using a cubic B-spline.
Towards the tip the bump placement and movement range
are increased in chord-wise position as the shock is sat
further downstream at the tip. The initial design used at the
start of the optimization process had bumps positioned
with approximately 60% of the bump downstream of the
datum shock, as is known to be beneficial from previous
work. The objective function for the optimization was blade
efficiency and the simulations were carried out at 98%
simulated choke. The optimizations were carried out on the
Rolls-Royce CFMS cluster using the MAM method. The Figure 10.29 Optimized R37 Bump
geometry of the optimized shock bump can be seen in (Blue) Added To The Datum Blade
Figure 10.28. A slice at 60% span is shown. The 3D Geometry (Grey)
geometry compared to the datum is shown in Figure
10.29.
The bump applied to the datum geometry varies radially, with the maximum bump amplitude and
273

width localized between 40 and 60% span. This makes sense as the strongest shock location, largest
separation and maximum adjoint sensitivity occur around mid-span for Rotor 37, and therefore
greater shock control is needed in this region. The resulting variation from hub to tip of the geometry
demonstrates the benefit provided by optimizing the geometry. Without optimization it would be
difficult to manually specify the bump position, width, amplitude and asymmetry, which would result
in reduced benefit.
10.6.13 Analysis of the R37 Optimized Bump Design
The flow features for the resulting, optimized, continuous bump design is compared to the datum in
Figure 10.30. The datum shock position is shown via a white line on the optimized geometry. It can
be seen how the use of bumps has delayed the shock. The reduction in separation for the optimized
design can be seen in Figure 10.32. The delay of the shock position has reduced the separation
initiation point and the volume of separated flow. The performance of this geometry is compared to
the best individual bumps geometry (not described in detail here) and the datum in Table 10.4. It
can be seen that the efficiency benefit is greatest for the continuous bump design. The efficiency is
increased by 1.48%, while the pressure ratio is also increased. A summary of previous optimization
results for Rotor 37 by various researchers is given by [ John et al.]356. The maximum efficiency
benefit achieved by those studies was around 1.7-1.9% (without decreasing PR). These optimizations
were able to modify parameters such as blade camber, thickness, lean and sweep though, so had

Figure 10.30 Datum (Left) And Optimized (Right) Rotor 37 Static Pressure Contours. Flow Direction
Right To Left – Courtesy of [John et al.]

greater design flexibility


than the current shaping PR Delta PR Efficiency / Delta efficiency
approach. This shows /% % /%
that the efficiency benefit Datum 2.05 - 85.45 -
provided through the Individual 2.06 0.51 86.21 0.76
application of shock Cont. 2.08 1.2 86.93 1.48
control bumps is
significant, considering Table 10.4 Rotor 37 Optimized Bump Performance Comparison –
Courtesy of [John et al.]
the only geometry

356John, A., Shahpar, S., and Qin, N., 2017. “Novel compressor blade shaping through a free-form method”.
Journal of Turbomachinery, 139(8).
274

change is the addition of bumps.


Figure 10.31 shows the passage
flow for the datum and optimized
geometries at 50% span. The
effect of the bump delaying the
shock can be seen, with the datum
shock position shown by the black
line. The shock has been delayed
by over 12% chord at this height.
Just upstream of the shock the
Mach number contour is lower,
suggesting pre-compression has
occurred. The boundary layer
separation that forms the wake,
highlighted by the dark blue, low
Figure 10.31 Datum (Left) And R37 Optimized (Right)
velocity region, has reduced in Flow Features At 50% Span – Courtesy of [John et al.]
width by 26% at the trailing edge
for the optimized design. Figure
10.33 shows the datum and optimized lift plots. It can be seen how the shock has been delayed. The
Cp increases just upstream of the shock, showing that the bump has carried out pre-compression.
The jump in pressure across the shock is also lower for the optimized design than for the datum,
indicating it has been weakened. Because the shock is delayed, it has become swallowed by the
passage, causing an acceleration near to the leading edge on the blade pressure surface. This can be
seen in the lower surface spike on the lift plot.
10.6.14 Performance Across the Characteristic for R37
The off-design performance is a key feature of blade aerodynamics. The characteristics for the datum
and optimized designs are shown in Figure 10.34. An efficiency and pressure ratio increase has
been achieved across the characteristic. The choke mass flow does not appear affected, although it is
possible that the choke margin has been modified at other rotor speeds due to the throat area being
reduced by the bump. The simulation results suggest a reduction in stall margin for the optimized
design. This is due to the shock bump being mis-placed at conditions away from where it was

Figure 10.32 Datum (Left) And Optimized (Right) Rotor 37 Separated Flow Contours (Orange). Flow
Direction Right To Left.
275

designed, leading to increased


separation and thus reduced stall
margin. This section has
demonstrated the benefit that can
be achieved by applying shock
control bumps to a compressor
blade without modifying the entire
blade geometry. This shows that a
significant benefit is possible
through geometry modifications
via bumps just in the shock region.
10.6.15 Conclusion
This work has demonstrated how
shock control bumps can be used
to improve the performance of
transonic fan/compressor blades.
Blade geometries that incorporate
Figure 10.33 Lift Plots For The Datum and Optimized
shock control bumps have the Geometries at 60% Span – Courtesy of [John et al.]
ability to reduce shock loss and
reduce/eliminate shock-
induced separation and
increase both efficiency
and stall margin. Shock
control bumps have the
benefit that only small
modifications to the blade
geometry are required to
achieve these
improvements, compared
to the large changes
required by blade designs
that make use of negative
camber or similar shock
control approaches. It has
been demonstrated that
both the efficiency and
pressure ratio of a highly
loaded compressor blade
can be increased across a
range of flow rates by
Figure 10.34 R37 Optimized Characteristic Vs Datum – Courtesy of
delaying the shock and [John et al.]
significantly reducing the
separation and wake. For a
modern fan blade the optimized bump design eliminated the majority of separation, reduced the
thickness of the wake and extended the stall margin. For further and complete info, please consult
the [ John et al.]357.

357John, A., Shahpar, S., and Qin, N., 2017. “Novel compressor blade shaping through a free-form method”.
Journal of Turbomachinery, 139(8).
276

10.7 Case Study 5 - 3D Multi-Objective Design Optimization of a Transonic


Compressor Rotor358
Citation : Ernesto Benini, Three-Dimensional Multi-Objective Design Optimization of a Transonic
Compressor Rotor, Journal of Propulsion and Power 2004 20:3, 559-565
A method for transonic compressor multi-objective design optimization was developed and applied
to the NASA Rotor 37, a test case representative of complex three-dimensional viscous flow
structures in transonic blading. The optimization problem considered was to maximize the isentropic
efficiency of the rotor and to maximize its pressure ratio at the design point, using a constraint on the
mass flow rate. The 3D Navier–Stokes code CFXTASCflow was used for the aerodynamic analysis of
®

blade designs. The capability of the code was validated by comparing the computed results to
experimental data available in the open literature from probe traverses up and downstream of the
rotor. A multi-objective evolutionary algorithm was used for handling the optimization problem
that makes use of Pareto optimality concepts and implements a novel genetic diversity evaluation
method to establish a criterion for fitness assignment. The optimal rotor configurations, which
correspond to the maximum pressure ratio and maximum efficiency, were obtained and compared
to the original design.
10.7.1 Introduction
In modern high-performance aircraft engines, compressor stages are required to operate with the
highest values of both the efficiency and the compression ratio. This helps minimize the fuel
consumption and decrease the engine weight and size due to the reduction in the number of stages
and their cross-sectional area. Transonic compressors have been developed during the past three
decades to achieve this goal. In these machines, the inner portion of the rotor blade operates at
subsonic relative speeds, whereas in the outer part the flow is supersonic. This is a direct
consequence of the high relative Mach numbers of the flow approaching the blade needed to obtain
a high-compression ratio per single stage.
The main drawback is that the rotors experience the existence of intense shock waves that are
generated close to blade tip and over part of the span. It is well known that shock waves are
responsible for high aerodynamic losses and entropy generation that negatively influence overall
efficiency. At the same time, however, the shock determines a significant increment in the blade
diffusion and, thus, contributes to achieving a high-compression ratio.
The presence of complex shock structures that interact with the main flow and wall boundary layers
explains why the design of such machines is so challenging. In fact, the design principle is not to avoid
shocks (as is the case in subsonic compressors), but to control their locations and strengths to achieve
maximum performance.
Computational fluid dynamics (CFD) is of practical importance for this purpose because it helps the
designer to better understand the features of the flowfield and establish the way the geometry of the
blade has to be designed and, if necessary, modified. Examples of CFD calculations on transonic
compressor blades can be easily found in the open literature 359-360. Navier–Stokes analyses, in
particular, make it possible to deal with the real three-dimensional design problem and to study the
effect of changes in the three-dimensional blade shape. These include both the application of high-
performance airfoils for transonic bladings, as well the use of sweep and lean in blade radial stacking

358 Ernesto Benini, “Three-Dimensional Multi-Objective Design Optimization of a Transonic Compressor Rotor”,
Journal of Propulsion and Power Vol. 20, No. 3, May–June 2004.
359 Miller,D. P., and Bryans, A. C., “The Relative Merits of Inviscid Euler3D and Quasi-3D Analysis for Design of

Transonic Rotors,” American Society of Mechanical Engineers, ASME Paper 88-GT-69, 1988.
360 2Hah, C., and Wennerstrom, A. J., “Three-Dimensional Flow Fields inside a Transonic Compressor with Swept

Blades,” Journal of Turbomachinery, Vol. 113, No. 1, 1991, pp. 241–251.


277

361-362.

CFD alone can be very useful in the framework of a conventional trial-and-error procedure of a
transonic blade, but it becomes even more powerful when combined with an optimization technique.
This aspect has recently been investigated by several researchers and the results documented in
relevant publications. [Lee and Kim]363 used a gradient-based method to optimize the blades of a
compressor using the three-dimensional Navier–Stokes equations.
[Ahn and Kim]364 developed an optimization technique for the NASA rotor 37 based on the response
surface method. [Oyama et al.]365 used evolutionary algorithms in the redesign of the blade of NASA
rotor 67. The latter was also adopted by [Tiow and Zangeneh]366 as a test case for the application of
a three-dimensional inverse design methodology. [Tiow et al.]367 developed and applied an inverse
design technique coupled with a simulated annealing optimization algorithm in the design of
transonic axial blade cascades.
In all of the cited works, optimization and inverse methodologies were adopted to minimize
aerodynamic losses. In the present work, a multi-objective optimization algorithm is applied with the
aim of optimizing a transonic rotor blade with respect to both the aerodynamic efficiency and the
pressure ratio. To achieve this goal, evolutionary algorithms16 are preferred to deterministic
algorithms in view of their ability to capture global optima and because of their intrinsic capability
to support an optimization problem involving multiple objectives. The paper is organized in the
following way:
First, a brief summary on the influence of three-dimensional blade geometry on performance of
transonic rotors is presented. This is done with the aim of giving a justification about the choice of
the design parameters. Then, the multi-objective optimization problem with regard to NASA rotor 37
is formulated, and the optimization tools used to deal with it are described. Finally, the results of the
optimization are presented and discussed.
10.7.2 Influence of Blade Shape on Transonic Compressor Flows
10.7.2.1 Effect of Blade Profiles
At high-subsonic relative Mach numbers, some areas of supersonic flow appear on both the suction
and the pressure surfaces. These areas are usually followed by shock waves, which interact with the
profile boundary layer and often lead to its separation. Since the development of the first and second
generation of controlled diffusion airfoils368, the knowledge regarding the design of shock-free or
shock-controlled profiles has dramatically improved, and this has led to successful increments in
stage efficiency, loading, and stall margin.
In blade profiles with a supersonic inlet flow, the greatest attention is directed toward the inlet region

361 Prince, D. C. J., “Three-Dimensional Shock Structures for Transonic/Supersonic Compressor Rotors,” Journal
of Aircraft, Vol. 17, No. 1, 1980, pp. 28–37.
362 Wennerstrom, A. J., “Experimental Study of a High-Through-Flow Transonic Axial Compressor Stage,” Journal

of Engineering for Gas Turbines and Power, Vol. 106, No. 3, 1984, pp. 552–560.
363 Lee, S. Y., and Kim, K. Y., “Design Optimization of Axial Flow Compressor Blades with Three-Dimensional

Navier–Stokes Solver,” American Society of Mechanical Engineers, ASME Paper 2000-GT-0488, 2000.
364 Ahn, C.-S., and Kim, K.-Y., “Aerodynamic Design Optimization of an Axial Compressor Rotor,” American

Society of Mechanical Engineers, ASME Paper GT-2002-30445, 2002.


365 Oyama, A., Liou, M.-S., and Obayashi, S., “Transonic Axial-Flow Blade Shape Optimization Using Evolutionary

Algorithm and Three-Dimensional Navier–Stokes Solver,” AIAA Paper 2002-5642, 2002.


366 Tiou, W. T., and Zangeneh, M., “Application of a Three-Dimensional Viscous Transonic Inverse Method to

NASA Rotor 67,” Journal of Power and Energy, Vol. 216, No. 3, 2002, pp. 243–255.
367 Tiou, W. T., Yiu, K. F. C., and Zangeneh, M., “Application of Simulated Annealing to Inverse Design of

Transonic Turbomachinery Cascades,” Journal of Power and Energy, Vol. 216, No. 1, 2002, pp. 59–73.
368 Hobbs, D. E., and Weingold, H. D., “Development of Controlled Diffusion Airfoils for Multistage Compressor

Applications,” Journal of Engineering for Gas Turbines and Power, Vol. 106, No. 1, 1984, pp. 271–278.
278

because it fixes the maximum mass flow capacity and produces the majority of the pressure rise 369.
Downstream of the inlet region, the flow is generally subsonic as a result of the presence of a passage
shock, and therefore, the criteria usually employed for subsonic blading may be appropriate there to
guide the designer toward achieving best performance. For a given operating point defined by the
inlet Mach number and incidence angle, the behavior of the inlet region is defined by the loading
distribution, particularly the position and magnitude of the peak load. Recently, this aspect has been
very well described by [Tiow et al.,]370 who demonstrated how a reduction in the peak load and the
shift in its position from the rear to the front result in diminishing the intensity of the shock patterns.
In the rear of the profile, the loading distribution is still regarded as very important and has to be
carefully specified371.
Some recent achievements in the field of transonic compressor bladings, however, have shown that
shock structure and strength cannot be controlled only by using a proper airfoil shape, but also by
taking advantage of the shape of the radial stacking curve, in particular including three-dimensional
lean and sweep.
10.7.2.2 Effect of 3-D Sweep and Lean
The concept of incorporating blade sweep for controlling shock structures and secondary flows has
been largely documented in the literature. In absence of shocks, sweep and lean are known to be
effective in reducing the onset and development of secondary flows within a blade row. [Yamaguchi
et al.]372 tested, in particular, the effect of forward sweep and described a qualitative mechanism for
suppressing secondary flows near the endwall region. They found that in forwardswept rotor blades
the accumulation of low-momentumfluid near the endwall tip region is much lower than in
conventional radial blades due to the decreasing radial migration of fluid particles within the
boundary layer, a phenomenon that follows the imbalance between the centrifugal force and the
pressure gradient.
In transonic compressors, the presence of shock structures further complicates the flow within the
blades. Since the experience of [Prince]373,6 it has been recognized that a three-dimensional-shaped
shock structure, as occurs in swept rotors, may be responsible for reduced shock losses. This would
be the consequence of the formation of a less strong oblique shock wave in the spanwise direction
compared to a high-strength normal shock, which is usually found in conventional transonic blades.
Recently, [Hah et al.]374 carried out an extensive experimental and numerical study on three transonic
rotors (unswept, aft swept and forward swept) and found that the forward-swept rotor had a higher
peak efficiency than the baseline unswept configuration, as well as 30% larger stall margin. However,
the aft-swept rotor showed almost the same peak efficiency as the baseline rotor, but a reduction of
about 40% in the stall margin.
The influence of blade lean in transonic compressor rotors has not been extensively described in the
literature. Results of both experimental and numerical studies refer, in fact, to the effect of lean (often

369 Cumspty, N. A., Compressor Aerodynamics, Longman Group UK, London, 1989.
370 Tiou, W. T., Yiu, K. F. C., and Zangeneh, M., “Application of Simulated Annealing to Inverse Design of
Transonic Turbomachinery Cascades,” Journal of Power and Energy, Vol. 216, No. 1, 2002, pp. 59–73.
371 Benini, E., and Toffolo, A., “Development of High-Performance Airfoils for Axial Flow Compressors Using

Evolutionary Computation,” Journal of Propulsion and Power, Vol. 18, No. 3, 2002, pp. 544–554.
372 Yamaguchi, N., Tominaga, T., Hattori, S., and Mitsubishi, T., “Secondary-Loss Reduction by Forward-Skewing

of Axial Compressor Rotor Blading,” Proceedings of 1991 Yokohama International Gas Turbine Congress, 1991.
373 Prince, D. C. J., “Three-Dimensional Shock Structures for Transonic/Supersonic Compressor Rotors,”

Journal of Aircraft, Vol. 17, No. 1, 1980, pp. 28–37.


374 Hah, C., Puterbaugh, S. L., and Wadia, A. R., “Control of Shock Structure and Secondary Flow Field Inside

Transonic Compressor Rotors Through Aerodynamic Sweep,” American Society of Mechanical Engineers, ASME
Paper 98-GT-561, 1998.
279

called dihedral) on subsonic linear cascades22 and subsonic annular cascades (bowed stators)375.
However, the use of lean is currently exploited in axial-flowfan rotors. The general guidelines that
result from these studies are difficult to apply to transonic rotors because of the strong interaction
between low-momentum fluid near the endwall and the shock waves. Recently, the use of negative
tip dihedral (where the profile stacking curve is skewed toward the direction of rotation) was
investigated numerically in a transonic rotor and the results showed a positive influence on the
overall rotor efficiency376.
10.7.3 Optimization of NASA Rotor 37
Rotor 37 designed at NASA John H. Glenn Research Center at Lewis Field, was selected for optimization
because it is a welldocumented test case, where strong interactions (such as corner stall,
shock/boundary-layer, tip–vortex, and tip leakage secondary interactions) occur. Even though it was
developed more than 20 years ago, rotor 37 is representative of design and performance levels
of the most advanced transonic
blades used today in gas turbine
inlet stages. Some design
information and overall stage
performance came from Reid and
[Moore]377, and detailed
measurement data were provided
by [Moore and Reid]. The main
dimensions of the rotor are given in
Figure 10.35.
The rotor has 36 multiple-circular-
src (MCA) blades, inlet hub–tip
diameter ratio of 0.7, blade aspect
ratio of 1.19, and a tip solidity of
1.29. The running tip clearance is
0.0356 cm (0.45% of the blade
span). The inner diameter increases
in the mean flow direction, whereas
the outer diameter decreases, and
the blades are stacked in the radial Figure 10.35 Meridional geometry of NASA rotor 37, from
direction in such a way that the flow AGARD
shows typical three-dimensional
features. Design performance at the nominal rotating speed of 17,188 rpm (at International
Standards Organization conditions) as estimated during the design computations are the following:
mass flow rate 20.19 kg/s, total pressure ratio 2.106, total temperature ratio 1.27, and adiabatic
efficiency 0.877. The experimental maximum mass flow rate at the choking condition is 20.93 kg/s.
The flow surveys were placed in stations 1 and 4 [Figure 10.36, (see378 Ref. 30)]. Cobra probes were
used for total pressure and flow angle measurements, with a thermocouple for total temperature
data; wedge probes were used for static pressure, and wall static pressure taps were employed on
the hub and tip walls.

375 Wisler, D. C., “Loss Reduction in Axial-Flow Compressors Through Low-Speed Model Testing,” Journal of
Engineering for Gas Turbines and Power, Vol. 107, No. 2, 1985, pp. 354–363.
376 Ahn, C.-S., and Kim, K.-Y., “Aerodynamic Design Optimization of an Axial Compressor Rotor,” American

Society of Mechanical Engineers, ASME Paper GT-2002-30445, 2002.


377 Reid, L., and Moore, R. D., “Design and Overall Performance of Four Highly Loaded, High-Speed Inlet Stages

for an Advanced High-Pressure-Ratio Core Compressor,” NASA TP 1337, 1978.


378 CFD Validation for Propulsion System Components,” AGARD-ar-355, AGARD, May 1998.
280

Figure 10.36 Measurement Stations within NASA Rotor 37, from AGARD

The purpose of the optimization considered herewas to maximize the two-objective function:

p04
𝐅 = (ηis , )
p01
Eq. 10.5
where ηis is the adiabatic efficiency and p04/p01 the total-tototal pressure ratio, both referred to
stations 1 and 4 of Figure 10.36. The optimization was conducted for one mass flow condition (m˙
/m˙ choke =0.98), this being one of the cases for which many of the experimental and computational
results are available to the public. The inlet total pressure and total temperature were fixed at p01
=101325 Pa and T01 =288.15 K. The optimization was carried out using an optimization method that
integrates a code for three-dimensional blade geometry parameterization, a Navier–Stokes solver,
and an optimization algorithm. A detailed description of this method, along with examples of its
application to design optimization of turbomachinery blades may be found elsewhere 379-380-381.
10.7.3.1 Blade Geometry Definition
To make the results of the optimization comparable to those regarding the baseline configuration,
the meridional contours of the hub and casing were not modified. Actually, a change in the meridional
area would have had a strong impact on the aerodynamic blockage and, therefore, on the compressor

379 Benini, E., and Toffolo, A., “Development of High-Performance Airfoils for Axial Flow Compressors Using
Evolutionary Computation,” Journal of Propulsion and Power, Vol. 18, No. 3, 2002, pp. 544–554.
380 Benini, E., “Optimal Navier Stokes Design of Compressor Impellers Using Evolutionary Computation,”

International Journal of Computational Fluid Dynamics (to be published).


381 Benini, E., and Toffolo, A., “A Parametric Method for Optimal Design of Two-Dimensional Cascades,” Journal

of Power and Energy, Vol. 215, No. A4, 2001, pp. 465–473.
281

flow capacity.
The rotor blade geometry was
parameterized using three profiles
along the span (hub, midspan, and tip
profiles), each of which was described
by camber and thickness distributions
(Figure 10.37). These were defined
by fourthorder Bezier polynomials,
where only the values of the ordinates
of the control points (θ for the camber
line and t the thickness) were allowed
to vary as independent design
variables.
The blade surfacewas then obtained
by the interpolation of profile
coordinates in the span direction by
use of spline curves. When a proper
value of the θ coordinate of the first
midspan and the tip profiles’ control
point with respect to the hub profile
were specified the effect of blade lean
was achieved. The use of blade sweep,
on the other hand, which could be
obtained in principle by giving a
different z coordinate to the three
profiles,was not investigated. The
results of preliminary calculations
indicated that, because the casing
contour is fixed here, a forward-swept
rotor would necessarily have a greater
diameter and, therefore, would lead to Figure 10.37 Parameterization of a Compressor Airfoil
higher aerodynamic losses caused by
higher incidence tip Mach numbers.
For the same reason, the chord distribution along the span was not changed compared to the original
design.
A total of 14 parameters for the camber lines plus 9 parameters for the thickness distributions (the
leading-edge and trailing-edge radii were not changed), that is, 23 parameters in total, were used to
describe the three-dimensional shape of the rotor. Each parameter was given a range of variation. To
avoid the creation of rotors having very different mass flow rates from that of the baseline
configuration, the code that handled the blade parameterization calculated the geometric throat area
between adjacent blades: Only the geometries that gave throat areas in the range of ±0.2% with
respect the original geometry were then simulated, and the others were disregarded and eliminated
before processing.
10.7.3.2 Flow Solver: Description and Validation
The CFD code CFX-TASCflow® was used to calculate the flowfield around the rotor, where the 3D
Reynolds averaged Navier–Stokes equations (RANS) are solved using a finite-elementbased finite
volume method. An algebraic multigrid method based on the additive correction multigrid strategy
was used along with the second order skew upwind differencing scheme with physical advection
correction.
282

The code was first validated against experimental data provided by [Moore and Reid]. Amultiblock
structured grid of about 240 K nodes per single passage (Figure 10.38) was adopted following the
guidelines provided in a recent AGARD report. The flow region close to the profile wall was
discretized using an
O-type grid, whereas
the outer part was
meshed using an H-
type grid. The effect of
the tip clearance was
also modeled. The κ-ε
turbulence model35
along with
standardwall
functions were
employed. Thewalls
were treated as
smooth and adiabatic.
The boundary
conditions were fixed
as follows: At the
domain inlet, the total
pressure, total Figure 10.38 Multiblock Grid Used in the Simulations
temperature, and flow
angle were imposed; at the outlet, average static pressure was applied for both near-stall and near-
choke conditions; periodic boundary conditions were imposed on the lateral faces of the flow domain.
An angular velocity corresponding to the nominal rotational speed was applied to the rotor. The rotor
performance was calculated
over the entire operating range.
For each simulation, the
convergence criterion was
established when normalized
rms residuals were less than
5×10−7. Each simulation took
about 4 hours to complete on a
Workstation AlphaServer ES40.
Results of the code validation are
given in [Benini] 382. With regard
to overall performance, both the
adiabatic efficiency and pressure
ratio were slightly
underestimated in all of the
operating conditions. In
particular, the pressure ratio Figure 10.39 Comparison Between Calculated and Experimental
seemed to have a dominant Mach Number Contours at 90% Span and m˙ /m˙ choke = 0.98
effect on the overall efficiency, as :Maximum Mach = 1.576, Minimum Mach = 0, and Contour Interval
= 0.031.
suggested by the radial plots,
based on pitch-averaged data at

382Ernesto Benini, “Three-Dimensional Multi-Objective Design Optimization of a Transonic Compressor Rotor”,


Journal of Propulsion and Power Vol. 20, No. 3, May–June 2004.
283

station 4 (see383 figure 5).


In fact, in a large portion
of the span the pressure
ratio was lower than the
measured one, and
because the temperature
ratio was predicted
rather well, a lower value
of the efficiency was
calculated. This tendency
seems to be in contrast
with other published
results regarding a
similar application of
CFX-TASCflow, where the
pressure ratio tended to
be overpredicted in all of
the conditions.
Figure 10.39 shows the
computational and
experimental results of
the relative Mach
number contour plots at
90% span and 98%
choke flow. The
calculated shock position
and resolution were
quite good as a result of
the alignment of the grid
lines along the mean flow
direction. The region
where the passage shock
and the boundary layer
on the suction side of an
adjacent blade intersect
was calculated quite well,
although the shock wave
front seemed wider than
the experimental one.
Hildebrandt, showed
that, to improve shock
resolution, a grid of about
500 K nodes, that is, finer Figure 10.40 Blade Geometries of the Baseline and Optimized
than the one used here, Configurations
should be used.
10.7.3.3 Multi-Objective Evolutionary Algorithm

383 See Above.


284

Evolutionary algorithms384 are optimization techniques that use both stochastic and deterministic
elements, where an artificial evolution process that imitates the natural evolution of biological
organisms is implemented. The evolution process starts with a randomly initialized population of
individuals (a set of points in the search space) that evolves following the Darwinian principle of the
survival of the fittest. According to this approach, new generations of solutions are created using
some simulated evolutionary operators, such as crossover and mutation: The probability of survival
for each individual depends on its fitness, that is, on how well it performs with respect to the
objective(s) of the optimization problem.
As a result, evolutionary algorithms are very well suited to deal with multi-objective problems
because they make use of an evolving population of solutions that is driven toward the set of the true
tradeoff among the objectives, the Pareto optimal set. For the same reason, they can be beneficially
applied to highly multidimensional problems, where an effective exploration of the design space can
be carried out only using populations with several individuals.
Here, the structure of the evolutionary algorithm follows the main steps of an evolution strategy and
has been described in detail in previous papers. After the parents are selected, reproduction using
uniform crossover and mutation are performed to generate offspring that are then evaluated. The
evaluation step includes reconstruction of rotor geometry from actual decision variables, CFD
analysis with postprocessing, and ranking according to the usual Pareto concepts385.
Then a genetic diversity evaluation method (GeDEM) is applied to establish a criterion for fitness
assignment and to build the next population of parents. In short, the GeDEM preserves genetic
diversity of the best-so-far population of candidate solutions to the optimization problem by
performing an additional evaluation after the common measure of objective fitness. This evaluation
ranks the solutions according to their fitness value and their reciprocal distance as a way to give more
reproduction chances to both highly fit and highly distant individuals. The loop starts again until the
predetermined number of generations has elapsed.
10.7.4 Results
During the optimization run, a population of 20 individuals evolved for a total of 100 generations.
Computations were performed in parallel on a four-processor Workstation AlphaServer ES40 and
the overall turn around time was about 2000 h. Because of time constraints, it was not possible to
continue the computations further. A mass flow rate boundary condition at the outlet was applied to
each configuration examined during the evolutionary process to fix the condition m˙/m˙ choke = 0.98.
Not shown here (see [Benini]386), the results of the optimization, that is, the performance of
optimized configurations with respect to both maximum efficiency and maximum pressure ratio.
These configurations correspond to the borders of the final Pareto front obtained after the
optimization. In fact, in the objective function space, the performance of the optimized individuals
was very much clustered around the original as a result of the tight constraint imposed on the mass
flowrate, which prevented the search algorithm from generating and evaluating geometries very
different from the original design.
At the chosen optimization point, an improvement in the adiabatic efficiency was achieved (+1.5%
with respect to the original geometry) without modifying the pressure ratio. On the other hand, an
optimal individual was obtained that showed a higher pressure ratio (+5.5%) with a slightly smaller
efficiency (−0.8%) compared to the original design. This behavior was observed over the entire
operating range of the compressor. In fact, for nearly the same pressure ratio, the efficiency-
optimized (E-O) rotor definitely showed a superior efficiency close to the choking condition (ηis =
+2%). Moreover, the operating range was very similar to the original one. The pressure ratio-

384 Deb, K., Multi-Objective Optimization Using Evolutionary Algorithms, Wiley, Chichester, England, U.K., 2001.
385 Deb, K., Multi-Objective Optimization Using Evolutionary Algorithms, Wiley, Chichester, England, U.K., 2001.
386 Ernesto Benini, “Three-Dimensional Multi-Objective Design Optimization of a Transonic Compressor Rotor”,

Journal of Propulsion and Power Vol. 20, No. 3, May–June 2004.


285

optimized (PR-O) rotor had instead a smaller operating range compared to the original one. The last
computation for which the CFD code was able to reach convergence occurred at m˙/m˙ choke = 0.951,
which perhaps denotes the stall limit.
The geometries of the optimized configurations are compared with the original design in Figure
10.40. The main differences could be found in both profile shape and radial stacking. Changes in the
profile shape concerned the region close to the trailing edge, as one might expect, because the
constraint on the geometrical
throat area led to cascade
geometries having very
similar inlet regions.
The E-O rotor had profiles
(particularly the one located
close to blade tip) with an
increased thickness toward
the rear and a different
curvature. Furthermore, the
blade profiles leaned
significantly in the direction
of rotation. This result
confirms the one obtained by
Ahn and Kim,12 even if the
amount of leaning was
considerably higher here
(+1.5 deg compared to +0.22
deg). The impact of this lean
on the structural strength of
the blade should, therefore,
be checked carefully. The PR-
O blade was characterized by
higher cambered profiles
toward the rear, with no
noticeable changes in the
thickness distribution
compared to the original
design. (An exception was the
profile at midspan, where the Figure 10.41 Mach Number Contours at 95 and 50 % Span of
maximum thickness is Baseline and Optimized Geometries
slightly lower). Again the
blade leaned substantially in the direction of rotation, even if in a less apparent way compared to the
E-O blade.
The effect of blade shape changes on rotor performance can be better understood by examining the
contours of the Mach number reported in Figure 10.41. From Figure 10.41, it appears that in the
E-O blade the shock intensity was reduced close to blade tip as the shock wave moved from a nearly
normal to a much more oblique pattern.
Actually, the normal shock wave at the tip was substituted by two oblique shocks of lesser intensity,
whereas at midspan the shock still weakened and became more oblique to the incoming flow. This is
also confirmed by Figure 10.42, where a comparison of the Mach number contours near the suction
surface for the three blades is given.
286

In the original configuration, a strong shock wave occurred within the blade passage that turned
normal to the casing and led to high aerodynamic losses and severe shock/boundary-layer
interaction.

(a) Base Line (b) Max. Effiiency

(c) Max. Pr. Ratio

Figure 10.42 Mach Number Contours on The Suction Surface of the Baseline and Optimized Blades

In both the E-O and PR-O blades, the shock bifurcated into two lesssevere branches and almost
vanished into two shock/boundary-layer interaction zones. In the E-O blade, this led to higher
efficiency, whereas in the PR-O blade the concomitant effect of an increased blade profile curvature,
which helped to achieve a high-pressure ratio, resulted in a more evident boundary-layer separation
toward the rear and, therefore, in a lesser overall efficiency. In addition, this is the reason why the
PR-O blade achieved a smaller operating range: The blade was no longer capable of effectively
withstanding flow deviation at reduced incidence angles without incurring in massive separations
(stall) toward the rear.
10.7.5 Conclusions
Amethod for three-dimensional multi-objective optimization of a transonic rotor blade was
developed and tested which was based on an evolutionary algorithm and a Navier–Stokes code. The
method was applied to the design optimization of NASA rotor 37 with the aim of achieving maximum
efficiency and maximum pressure ratio with a constraint on the mass flow rate. The rotor blade was
287

described using three profiles along the span, each of which was defined using parametric curves.
The effect of blade lean was considered by changing the mutual tangential coordinates of the three
profiles.
The optimization run was carried out on a multi-processor computer and demonstrated that the
overall adiabatic efficiency can be improved by approximately 1.5% (without changing the pressure
ratio in a significant way) by giving the blade a proper lean toward the direction of rotation and by
slightly changing the profile shape, especially toward the tip. This improvement followed from a
drastic modification in the shock structure within the blade passage.
The results also showed that the improvement in the overall efficiency, achieved in one operating
point, is maintained at off-design conditions. The results also showed that the pressure ratio can be
improved by about 5.5% by paying for a small efficiency drop (−0.8 %). This was achieved by leaning
the blade in the direction of rotation and by slightly increasing the profile curvature toward the rear
to assure a subsonic diffusion. In this case, however, the presence of a shock wave, although less
intense, accentuated the interaction between the shock and the boundary layer on the rear of the
suction surface, a phenomenon that possibly determined a reduction in the operating range of the
compressor.

10.8 Case Study 6 - Effect of Twist of a Wide-Chord Fan-Blade on the Aerodynamic


Performance of the Fan of a High-Bypass Turbofan Engine
Authors : Tariq Alsawy and Ahmed Farouk AbdelGawad
Appeared in : ICFD13-EG-6S01
Source : Proceedings of ICFD13: 13th International Conference of Fluid Dynamics, December, 2018.
Even a small change in the complex geometry of the wide-chord blade of a high-bypass turbofan
engine can completely change the performance of the fan. Deep research and continuous
investigation are required to understand the aerothermodynamics of this complex flow. This paper
aims to understand the influence of the wide-chord blade twist on the performance of a high-
bypass turbofan fan in take-off conditions. Blade twist of 45°, 55°, 65°, and 75° are studied 387.
The 65° blade twist is the baseline case. It was found that, the baseline case of 65° twist has wider
range of operation and high efficiency in high mass flow operating conditions due to the passage
shock waves that eliminate blade stall. Increasing the blade twist to 75° reduces the operating limits
of the fan. However, it has the highest efficiency in all the studied cases at low mass flow-rate
conditions. Reducing the blade twist to 55° increases the maximum obtainable mass flow and
pressure ratio. But this case has a smaller operating range compared to the baseline case. Further,
reducing the blade twist to 45° causes further decrease in the operating limits of the fan and
efficiency.
10.8.1 Introduction
The aviation industry consumes roughly 1.5 billion barrels of jet fuel each year, producing over 700
million tons of CO2 as well as contributing to roughly 30% of airlines direct operating costs.
Improving fuel efficiency of aircraft engines requires deep research and continuous investigation
aiming to understand the aero-thermodynamics of such complex flow [1]. From aero-
thermodynamics point of view, the geometry of every component, in the airflow direction, in the
aircraft engine must be tailored to offer the highest possible engine efficiency. Hence, the airflow in
the aircraft engine is very sensitive to both blade geometry and casing shape.
Many efforts were carried out to understand the influence of different parameters on aircraft engine
performance. [Amano and Xu] investigated the sweep effect on a transonic compressor blade. They

387Tariq Alsawy and Ahmed Farouk AbdelGawad, “Effect of Twist of a Wide-Chord Fan-Blade on the
Aerodynamic Performance of the Fan of a High-Bypass Turbofan Engine”, ICFD13-EG-6S01.
288

showed that both forward and backward sweeps have different effects on the secondary flow pattern
and shock wave structures, depending on the baseline case. [El–Sayed and Ibrahim] studied the effect
of different tip clearance shape on performance of an axial-flow compressor. They found that
decreasing tip clearance positively affects engine performance and surge margin.
Moreover, they found that linear-expanding gap leads to increase in both pressure ratio and
efficiency of the compressor stage compared with the linear-shrinking gap. [Banks et al.] investigated
the effect of tip clearance on the performance of a high-pressure shrouded turbine rotor blade. They
stated that shroud significantly affects the blade-tip flow field. Moreover, the integrated shroud
increases the turbine stage efficiency and decreases its sensitivity to clearance. [Karrabi and
Rezasoltani] investigated the effect of lean, twist, and bow of the blade on the performance of axial
turbines. Their results showed that turbine blade-twist clearly affects the performance, while sweep
and bow have minor effects on the performance.
However, there is little information in the literature regarding the three-dimensional flow features
of fan blades, especially, the studies that involve the influence of blade geometry on the fan
performance. [Frohnapfel et al.] experimentally measured the response of a fan rotor operating with
continuous inlet swirl distortion. Worst case scenario flow pattern associated with aggressive
aircraft maneuvers was tested. They demonstrated that the swirl distortion continues after
interaction with the fan
rotor which operates the
downstream cascade on
their off-design conditions.
[Han et al.] studied the
effect of casing shape on a
transonic fan performance.
It was found that the
shrinking case improves
the peak efficiency.
Convex shrinking case
reduces losses in tip area
for all operating
conditions, whilst concave
shrinking case reduces
losses in tip area at large
mass flow points and has
the opposite effect at small
mass flow points. [Hisse et
al.] numerically simulated
the shock wave
propagation in turbofan
engines intakes. [Hall et al.]
used a genetic algorithm
and computational fluid
dynamics to optimize the
design of the turbofan inlet
duct. [Chai et al.] studied
the tip clearance of wide-
chord fan-blade of a high-
bypass turbofan engine. Figure 10.43 Different Cases of Blade Twist
Their results indicated
that, after certain fan
289

angular acceleration, blade deformation occurs, which can cause sever rubbing or impact between
blade tip and the casing.
Less research investigations are available regarding the geometry of the modern wide-chord fan
blade and its effect on fan performance.
Thus, this paper presents a 3D transient CFD study of a 3D modelled fan of a modern high-bypass
turbofan engine, Rolls Royce Trent XWB, rotating at its reference 100% speed (2700 rpm) at take-off
conditions [11]. The parameter of interest of this paper is the twist of the fan-blade and its effect on
the fan aerodynamics and performance which directly affect the engine performance.
10.8.2 Numerical Models and Technique
10.8.2.1 Model Construction
A CAD model of a wide-chord fan of a modern high-bypass turbofan engine, Rolls-Royce Trent XWB,
is constructed using SOLIDWORKS software. The CAD geometry is constructed similar to the realistic
3D model of Trent XWB engine released by Rolls-Royce plc [12]. Main dimensions of the blade (tip,
root and blade sections necessary to build the model) were obtained from the model using “ImageJ”
software. The profile of the constructed blade was approximated to match the realistic 3D model
using multiple blade sections between the root and tip of the constructed blade.
The different studied cases of blade twist are shown in Figure 10.43. The approximated CAD model
used in this work is shown in Figure 10.43c. It has a blade tip angle of 65° and will be used as a
baseline case for this work. Varying the twist, for the studied cases, is done by changing the tip blade
angle, while holding the blade chord length and the
general blade shape constant for all cases. At all blade Number of blades 22
sections, the blade angles are set as a ratio of the tip angle Fan diameter 3m
at the baseline case (65°). Fan speed 2700 rpm
Changing the tip angle by certain ratio, will change all Root chord length 0.47 m
blade section by the same ratio keeping the general blade Tip chord length 0.64 m
shape approximately the same. Using this method results
in isolating the blade twist from other geometric Table 10.5 Approximated model
confounding factors. So, the results obtained from the dimensions and specifications
study for different cases are proportional to the blade
twist only. Main parameters of the approximated blade model is shown in Table 10.5. In this work,
in addition to the baseline case studying, blade twist angles of 45°, 55°, and 75° are studied as shown
in Figure 10.43a, Figure 10.43b, and Figure 10.43d for the 45°, 55°, 75° twist, respectively.
10.8.2.2 Domain Construction and Meshing
Due to the similarity of the fan blades and to save computational effort and time, only one blade is
considered. ANSYS DesignModeler© is used to build the flow domain for a single blade passage. This
is achieved by implementing FlowPath and ExportPoints features in ANSYS DesignModeler. These
features allow the domain to be exported to TurboGrid to generate the computational grid (mesh). A
single blade passage with a structured H-type grid is generated by TurboGrid. Number of elements,
which is held constant for all cases, is 0.6 million elements for single blade passage. Mesh for the
baseline case is shown in Figure 10.44. As it is clear from Figure 10.44 b, careful consideration
was paid to achieve good meshing, with the finest possible resolution, for accurate results. The
technique of periodic domain is applied with the appropriate inlet and outlet surfaces, Figure 10.44
c, Figure 10.44 d.
10.8.2.3 Numerical Scheme
The CFD solver used to obtain the unsteady RANS solution is ANSYS FLUENT [3] [9] [13] [14]. Due to
the high rotational speed (2700 rpm) of the fan in this work, air is modelled as compressible low. k-ε
model is widely used for turbulence modelling in the numerical analyses of turbomachines [3] [7]
[15] [14]. Thus, for this study, the Realizable k-ε turbulence model is used. Realizable k-ε model
290

outperform both RNG and standard k-ε models in this type of numerical analysis as it provides
superior performance for flows involving round jets and flows under strong adverse pressure
gradients, separation, and recirculation [16]. Enhanced wall treatment is chosen to be the near-wall
model as it combines both the two-layer model with enhanced wall functions. Compressibility effects
correction is added to the used turbulence model.
Sliding mesh technique is used to simulate the rotation of the fan blade. This is mainly because sliding
mesh provides relatively accurate solution compared to other methods such as MRF (Multiple
Reference Frame) [13] [16]. Only one mesh zone exists in this work, which is a rotating mesh zone.
The rotational speed of this zone is set to be the same reference 100% speed of the Trent XWB which
is 2700 rpm.
A Pressure Inlet boundary condition is used at the inlet, where the total pressure is set to be
atmospheric pressure and total temperature is assumed to be 300 K at take-off conditions. As an
initial value, static pressure at inlet is set slightly less than atmospheric pressure to provide more

(a) General view of the (b) Zoomed-in mesh at the


meshing of the blades blade root

(c) Periodic domain isometric view, (d) Periodic domain meridional view,
showing inlet (green) and,outlet (red). showing inlet at left and outlet at right

Figure 10.44 Mesh Generated by TURBOGRID for the Baseline Case


291

stable initialization for the solution. The outlet of the domain is set to be Pressure Outlet which
requires static pressure input. For every studied case, the static pressure of the outlet is changed to
obtain different operating point for the blade. Both blade and hub are set as a moving wall with a
speed of 0.0 rpm relative to the rotating domain. Shroud wall is set to be a stationary wall, which act
as a counter rotating wall to the rotating domain. Rotating periodic boundary condition is applied at
both sides of the rotating zone.
10.8.2.4 Methodology and Time Stepping
Momentum, continuity, turbulence, and energy equations are discretized using Second-Order
scheme. However, Transient formulation is discretized using Bounded Second-Order scheme, as this
scheme provides the same accuracy of the second-order, but it is more stable [16]. To ensure stable
and accurate solution for this kind of complex flow domain, solution is done over gradual steps. All
equations are set to First-Order discretization scheme. Solution is initialized using a low speed (100
rpm) and gradually increased to the full speed (2700 rpm). When reaching the full speed,
discretization schemes are gradually changed from First to Second-Order.
Time step size is chosen to be small enough to attain acceptable accuracy while consuming
reasonable computational time. Time step size is set to be 2.5×10-5 s. So that, the rotor blade crosses
full blade passage in approximately 40 timesteps and perform full revolution in approximately 880
timesteps [14].
Solution monitors are monitoring total pressure and mass flow rate at inlet and outlet boundaries of
the domain. Solution is aborted when monitored variable reaches to an approximately constant value
and becomes approximately time independent. If, for different operating conditions, solution
fluctuation starts to occur in a small amplitude, the solution for this point is taken as the mean of this
fluctuation. Moreover, the point is marked as a limit of the operating range of the tested case on which
the blade starts to stall.
10.8.3 Numerical Results
10.8.3.1 Performance of All Cases
The isentropic efficiency is defined as the isentropic compression work divided by actual total work
and is calculated by [7]:
𝛾−1
p 𝛾
(p02 ) −1
01
η=
T
(T02 ) − 1
01
Eq. 10.6
All the obtained results are for one blade passage. Starting from case of 65o twist as a baseline case,
the performance of all the four cases is shown in Figure 10.45. It can be seen in Figure 10.45 a that
increasing the twist angle to 75° decreases both mass flow rate and pressure ratio limits of the fan.
However, decreasing the twist angle to 55° has positive effect on both mass flow rate and pressure
ratio. The fan blade achieves higher pressure ratio at higher mass flow rate before the fan blade
passage chokes. Nevertheless, the fan blade in this case has less operating range before the blade
stalls. Further decrease in the twist angle to 45° negatively affects the performance of the fan blade.
Moreover, in this case, the fan blade has more restricted operating range than 55° twist case.
Figure 10.45 b shows that, although the blade twist of 75° case has less mass flow rate and pressure
ratio limits, it is the most efficient case to operate at low mass flow rate conditions (54.5 kg/s)
compared to the other cases. The baseline case of 65° twist has the highest peak efficiency among the
tested cases when operating at higher mass flow rate conditions (65 kg/s) and pressure ratio of (1.4).
The 55° blade twist case has slightly less peak efficiency than the 65° twist case. However, the fan
292

blade in this 55° twist case can deliver more mass flow rate (67 kg/s) at higher pressure ratio (1.45).
The 45° twist case shows very poor efficiency compared to other cases
Further analysis is delivered in this paper to the flow field for the cases operating at high mass flow
rate conditions (blade twist of baseline case, 55° twist and 45° twist) at their highest efficiency. The
flow field for the cases operating in low mass flow rate conditions are analyzed at their highest
efficiency as well. Analyzed points are marked in blue color in Figure 10.45 b for high mass flow
cases and red color for low mass flow cases. The flow field analysis aims to investigate the reason
why some cases outperform the other cases in almost the same operating conditions.

(a) Total (b) Isentropic


pressure ratio efficiency

Figure 10.45 Performance Curves for the Studied Cases

10.8.3.2 Blade Twist Effect at High Mass Flow Rate Conditions


The effect of changing twist angle is more significant near the tip of the blade. Mach number
distribution of the flow field at 85% blade span is shown in Figure 10.46 for high mass flow rate
points mentioned before. Mach number distribution in Figure 10.46a for the baseline case of 65°
twist looks like the Mach number distribution obtained in [7] for the high mass flow points. However,
in the present work, there is significant difference between cases because of the change in blade
geometry. It can be seen in Figure 10.46 that Mach number upstream and downstream the leading-
edge shock waves varies significantly for the different cases, so as the shock strength and shock
losses.
Because of the blunt leading edge of the blade, a detached bow shock is formed upstream the leading
edge. However, further traveling into the suction surface, the bow shock turns to an oblique shock.
The downstream Mach number of the oblique shock is supersonic, which is observed in Figure
10.46. For pressure surface of the blade, a normal shock wave is formed in the passage with
downstream Mach number being subsonic.
In Figure 10.46 b, and Figure 10.46 c the suction-side shock/boundary layer interaction leads to
blade stall and the boundary layer rapidly transition to turbulence. Similar flow dynamics at
transonic fan blade section is observed in [17]. The pressure side has attached turbulent boundary
layer which has less losses than the suction side separated boundary layer. When the suction-
side/boundary layer interaction moves closer to the leading edge, the rapid growth of the suction-
side boundary layer and flow separation can choke the flow further downstream leading to shock
293

waves impinging on the leading edge of the pressure surface [17]. This can be clearly observed in of
Figure 10.46 b and Figure 10.46 c.
For the 65° twist case shown in Figure 10.46 a, Mach number upstream the bow shock is the lowest
among the other cases, leading to reduced shock strength and shock losses. Shock waves cover the
entire suction surface, increasing the pressure on this surface and eliminating blade stall. This
explains the high efficiency of this case in these conditions. Further stretching in the passage, a
normal shock wave is developed due to the supersonic Mach number downstream the oblique shock.
The normal shock impinges on the pressure side of the blade leading to slight boundary layer
separation which reattaches to the blade before the trailing edge.
Figure 10.46 b shows Mach number distribution for the 55° twist case. A bow shock is developed
upstream the leading edge. Stretching downstream, the bow shock attaches and results in an attached
oblique shock. The downstream Mach number of the oblique shock is supersonic resulting in
developing a normal shock in the passage impinging on the suction surface.
The Mach number downstream the oblique shock is higher than the 65° twist case, which in turn
increases the shock strength and shock losses. The suction side/boundary layer interaction is more
obvious in this case. The normal shockwave impinging on the leading edge of the blade results in
growth and separation of the boundary layer. Combining the boundary layer separation with the
increase in the upstream Mach number can explain the decreased efficiency of this case compared to
the 65° twist case.

(a) Baseline case 65° twist (b) Case of 55° twist

(c) Case of 45° twist

Figure 10.46 Mach Number Distribution for High Mass Flow Point

Figure 10.46c shows the Mach number distribution for the 45° twist case. The bow shock developed
upstream the leading edge turns to oblique shock on the leading edge of the blade. Mach number
downstream the oblique shock is the highest for the different cases. The suction surface
294

shock/boundary layer interaction


leads to severe boundary layers
growth and separation. The separated
flow reduces the area available for the
flow in the passage leading to flow
acceleration and further development
of normal shock waves in the passage
[17, 18]. Combining all these effects
interprets the low efficiency of this
case.
Consequently, the 65° case has less
losses compared to the 55° twist case.
But the 55° twist case has higher flow
speed leaving the blade (increasing
mass flow rate and pressure ratio).
The 45° twist case has the highest
losses. This agrees with Figure 10.47
that shows the total pressure over
blade span at the domain outlet. The Figure 10.47 Total Pressure Over Span for High Mass Flow
55° twist case has higher total points at domain outlet
pressure relative to 65° twist case
because of the faster flow leaving the
blade as noted in Figure 10.46 b. The 45° twist case has large total pressure drop at 80% span
which reduces the overall total pressure at the blade outlet and in turn reduces efficiency.

(a) Baseline case 65° twist (b) Case of 55° twist

(c) Case of 45° twist

Figure 10.48 Entropy distribution for high mass flow points


295

Figure 10.47 shows the variation of the total pressure along the blade span for high mass flow
points of different twist cases at domain outlet. Flow field entropy is shown in Figure 10.48 at 85%
span for 65°, 55°, and 45° twist cases, respectively. It can be noticed that the 65° twist case has lower
entropy values than the 55° twist case. This agrees with the slight increase in efficiency of the 65°
twist case. The shock losses are observed in all cases in Figure 10.48. However, entropy is
significantly raised by the boundary layer separation more than the shock losses. Accordingly, blade
stall is the dominant factor in reducing effeminacy in Figure 10.48 c of the 45° twist case that has
severe entropy rise due to the blade stall. This further agrees with the decreased efficiency of this
case.
10.8.3.3 Blade Twist Effect at Low Mass Flow Rate Conditions
At low mass flow rate points, Figure 10.49 shows Mach number distribution of the flow field
representing the baseline case of 65° twist and the 75° twist case, respectively. Figure 10.49 a
shows the Mach number distribution for the 65o baseline case. A bow shock upstream the leading
edge is developed and becomes closer to normal shock. Downstream the bow shock and stretching
into the suction surface, an oblique shock wave is developed. For this case, there is no obvious normal
shock developed in the passage compared to the large mass flow point, which may reduce the shock
losses. [Han et al.] obtained similar shock dynamics in the low mass flow point study. However, in the
present work, the suction-side boundary layer is separated due to the interaction with the suction-
side shock wave, which in turn leads to blade stall and efficiency decrease.

Figure 10.49 Mach number distribution for low mass flow points

Figure 10.49 b represents the Mach number distribution for the 75° twist case. A bow shock forms
upstream the leading edge. Stretching downstream, the bow shock turns to an oblique shock. The
Mach number downstream the oblique shock is supersonic, leading to a normal shock occurrence in
the passage. However, shock waves cover the entire suction surface of the blade which diminishes
the suction surface stall.
Eliminating the suction surface stall is key to reduce losses and increase efficiency. This can interpret
the increased efficiency for this case in the low mass flow conditions. Both the boundary layer of the
pressure surface and the normal shock interact leading to growth and detachment of the boundary
layer, which reattaches again due to the increased pressure on this surface.
The flow field entropy distribution is shown in Figure 10.50. The baseline case in Figure 10.50 a
shows increased entropy due to the suction-surface shock/boundary layer interaction leading to
blade stall. This explains the low efficiency of the baseline case in low mass flow rate conditions.
Figure 10.50 b shows the entropy distribution for the 75° twist case. There is far less entropy
296

increase in the flow field, which mainly because of the attached boundary layer due to shock waves
on the suction surface as mentioned before.
10.8.4 Conclusion
A three-dimensional unsteady
numerical analysis is used to analyze
the influence of a transonic fan blade
twist on the fan performance. A
modern fan blade of a high-bypass
turbofan engine “Rolls-Royce Trent
XWB” is approximately modelled to
perform the analysis. Twist of the
blade is changed by changing the tip
blade angle. Studied cases are a
baseline case of 65° twist, 45°, 55°,
and 75° twist. The rotating speed is
held constant at 2700 rpm through
all cases. One blade passage is
studied with rotating periodic
boundary condition. Performance
parameters are obtained for one
blade passage. The major findings
are as follows:
(1) Generally, the present
computational results compare well
with the results of others with
consideration of the operating
differences.
(2) The baseline case of 65° twist has
wider range of operation. Mass flow
ranges from 53 to 67 kg/s. Pressure
ratio ranges from 1.45 to 1.225,
respectively with mass flow points.
The best efficiency point is at 65 kg/s
and pressure ratio of 1.34. For this
case the increased efficiency is due Figure 10.50 Entropy Distribution for Low Mass Flow Points
to shock waves in the passage
covering the entire suction surface
eliminating blade stall.
(3) Increasing the blade twist to 75° reduces the operating limits of the fan. Mass flow ranges from
51 to 59.5 kg/s. Pressure ratio ranges from 1.35 to 1.175. However, it has the highest efficiency of the
studied cases while operating in low mass flow rate conditions (at mass flow of 55 kg/s and 1.35
pressure ratio). This is mainly because the shock waves cover the entire suction surface and diminish
the suction surface boundary layer separation. While the baseline case has slight boundary layer
growth and separation due to the suction surface shock/ boundary layer interaction.
(4) Reducing the blade twist to 55° increases the maximum mass flow limit to 68.2 kg/s. However,
the minimum mass flow before blade stall is reduced to 61 Kg/s, which reduces the overall
performance range of the fan. However, the blade has higher total pressure ratio. In high mass flow
operating conditions, this case has slight blade stall due to suction surface shock/boundary layer
297

interaction which slightly reduces efficiency compared to the baseline case. But it produces higher
mass flow and total pressure at its best efficiency point (mass flow of 68 kg/s and pressure ratio of
1.44).
(5) Further reducing the blade twist to 45° results in more decrease of the operating limits of the fan.
Moreover, it greatly decreases the efficiency due to increased shock waves strength and blade stall
driving this case out of competence with other cases.
(6) According to this blade design, for low mass flow operating conditions (54 to 58 kg/s), the 75°
blade twist is more efficient. At high mass flow operating conditions (65 to 68 kg/s), both 55° and 65°
blade twist are more efficient. The 65° twist case has wider operating range before the blade stalls.
However, the 55° twist case can produce more mass flow rate and pressure ratio but with limited
operating range before the blade stalls. The 45° twist case stalls, at all operating points, results in a
greatly decreased efficiency compared to other cases.
10.8.5 References
[1] R. Bhaskaran, T. Wood, U. Paliath, and A. Breeze-String fellow, “Towards Large Eddy Simulation
of a 3D Transonic Fan,” 46th AIAA Fluid Dynamics Conference, Washington, D.C., USA, 2016.
[2] R. S. Amano, and C. Xu, “Blade Sweep Effects of Turbomachinery,” 43rd AIAA Aerospace Sciences
Meeting and Exhibit, Nevada, USA, 2005.
[3] A. A. A. El-Sayed, and M. M. Ibrahim, “Numerical Investigation of Different Tip Clearance Shape
Effects on Performance of an Axial Flow Compressor Stage”, The Online Journal on Power and Energy
Engineering (OJPEE), Vol. 1, No. 2, 2010.
[4] W. V. Banks, A. A. Ameri, and J. P. Bons, “Numerical Investigation of Effects of Tip Clearance in
Shrouded Turbine Rotor Blade Performance,” 2018 AIAA Joint Propulsion Conference, Cincinnati,
Ohio, USA, 2018.
[5] H. Karrabi, and M. Rezasoltani, “The Effect of Blade Lean, Twist and Bow on the Performance of
Axial Turbine at Design Point,” ASME 2011 International Mechanical Engineering Congress and
Exposition, Colorado, USA, 2011.
[6] D. J. Frohnapfel, W. F. O’Brien, and K. T. Lowe, “Fan Rotor Flow Measurements in a Turbofan Engine
operating with Inlet Swirl Distortion,” 55th AIAA Aerospace Sciences Meeting, Texas, USA, 2017.
[7] L. Han, Y. Wang, X. Zhang, X. Zhang, and D. Wei, “Effect of Casing Shape Surrounded Rotor on
Aerodynamic Performance of a Transonic Fan,” 53rd AIAA/SAE/ASEE Joint Propulsion Conference,
Atlanta, Georgia, USA, 2017.
[8] J. Thisse, C. Polacsek, and J. Mayeur, “Numerical Simulations of Shock-Wave Propagation in
Turbofan Intakes,” 22nd AIAA/CEAS Aeroacoustics Conference, Lyon, France, 2016.
[9] Z. M. Hall, V. Ahuja, R. Hartfield, A. Shelton, and A. Ahmed, “Optimization of a Turbofan Inlet Duct
using a Genetic Algorithm and CFD,” 27th AIAA Applied Aerodynamics Conference, USA, 2009.
[10] X. Chai, P. Han, T. Shi, and Z. Wang, “The Study of Tip Clearance of a Wide-Chord Fan Blade of a
High Bypass Ratio Turbo-Fan Engine,” ASME Turbo Expo 2014: Turbine Technical Conference and
Exposition, Düsseldorf, Germany, 2014.
[11] “EASA.E.111 Rolls-Royce plc Trent XWB Series Engines,” European Aviation Safety Agency, 2018.
[12] “World's Most Efficient Large Aero-Engine,” Rolls-Royce plc, 2018. [Online]. Available:
www.rollsroyce.com/products-and-services/civilaerospace/airlines/trent-xwb.aspx.
[13] P. Gullberg, and R. Sengupta, “Axial Fan Performance Predictions in CFD, Comparison of MRF and
Sliding Mesh with Experiments,” SAE International, 2011.
[14] D. Brzozowski, O. Uzol, Y.-C. Chow, J. Katz, and C. Meneveau, “A Comparison of Unsteady RANS
Simulations with PIV Data in an Axial Turbomachine,” ASME Fluids Engineering Division Summer
Meeting and Exhibiti, Houston, TX, USA, 2005.
[15] J. Åhman, CFD Validation of Tip Clearance Flows in an Axial Compressor, Luleå, Sweden: Luleå
University of Technology, 2016.
[16] “ANSYS Fluent User's Guide Release 15.0,” Canonsburg, Pennsylvania, USA, ANSYS, Inc, 2013.
298

[17] P. K. Ray, and W. N. Dawes, “Detached-Eddy Simulation of Transonic Flow Past a Fan-Blade
Section,” 15th AIAA/CEAS Aeroacoustics Conference (30th AIAA Aeroacoustics Conference), Miami,
Florida, USA, 2009.
[18] S.-M. Li, and R. Ramakrishnan, “A CFD Study and Performance Evaluation of Service-Run Variable
Vanes in a High-Pressure Compressor of a Turbofan Engine,” 54th AIAA Aerospace Sciences Meeting,
San Diego, California, USA, 2011.
299

11 Radial Flow
Up to now we were mostly concern with Axial flows. Now we pay homage to Radial flows which
designed in many everyday life tools. According to dictionary, in radial flows, the working fluid is
flowing mainly along the radii of rotation.

11.1 Centrifugal Compressor


Centrifugal compressors, as depicted in Figure
11.1, sometimes termed Radial compressors,
are a sub-class of dynamic axisymmetric work-
absorbing turbomachinery388. The idealized
compressive dynamic turbo-machine achieves a
pressure rise by adding kinetic energy/velocity
to a continuous flow of fluid through the rotor or
impeller. This kinetic energy is then converted
to an increase in potential energy/static
pressure by slowing the flow through a diffuser.
The pressure rise in impeller is in most cases
almost equal to the rise in the diffuser section.
11.1.1 Operation Theory
In the case of where flow simply passes through
a straight pipe to enter a centrifugal
compressor; the flow is straight, uniform and
has no vorticity. As illustrated below α1 = 0°. As
Figure 11.1 Centrifugal impeller with a highly
the flow continues to pass into and through the polished surface likely to improve performance
centrifugal impeller, the impeller forces the flow
to spin faster and faster. According to a form of
Euler's fluid dynamics equation, known as pump and turbine equation, the energy input to the fluid
is proportional to the flow's local spinning velocity multiplied by the local impeller tangential
velocity. In many cases the flow leaving centrifugal impeller is near the speed of sound (340 m/s).
The flow then typically flows through a stationary compressor causing it to decelerate. These
stationary compressors are actually static guide vanes where energy transformation takes place. As
described in Bernoulli's principle, this reduction in velocity causes the pressure to rise leading to a
compressed fluid.
11.1.2 Similarities to Axial Compressor
Centrifugal compressors are similar to axial compressors in that they are rotating airfoil based
compressors as shown in the adjacent figure.389,390 It should not be surprising that the first part of
the centrifugal impeller looks very similar to an axial compressor. This first part of the centrifugal
impeller is also termed an inducer. Centrifugal compressors differ from axials as they use a greater

388 Shepard, Dennis G. “Principles of Turbomachinery”, McMillan. ISBN 0-471-85546-4. LCCN 56002849, 1956.
389 Lakshminarayana, B. (1996). “Fluid Dynamics and Heat Transfer of Turbomachinery”, New York: John Wiley
& Sons Inc. ISBN 0-471-85546-4.
390 Japikse, David & Baines, Nicholas C., “Introduction to Turbomachinery”, Oxford: Oxford University press.

ISBN 0-933283-10-5, 1997.


300

change in radius from inlet to exit of the rotor/impeller. The 1940s-era German Heinkel HeS 011
experimental aviation turbojet engine was the first aviation turbojet design to have any sort of
"mixed compressor" design in its fore-sections, as it had a single-stage "diagonal flow" main
compressor ahead of a triple-stage axial unit, driven by a twin-stage turbine.
11.1.3 Components of a simple Centrifugal Compressor
A simple centrifugal compressor has four components: inlet, impeller/rotor, diffuser, and collector.
Figure 11.2 shows each of the components of the flow path, with the flow (working gas) entering
the centrifugal impeller axially from right to left (blue). As a result of the impeller rotating clockwise
when looking downstream into the compressor, the flow will pass through the volute's discharge

Figure 11.2 Cut-Away View of a Turbocharger showing the Centrifugal Compressor

cone moving away from the figure's viewer. The inlet to a centrifugal compressor is typically a simple
pipe. It may include features such as a valve, stationary vanes/airfoils (used to help swirl the flow)
and both pressure and temperature instrumentation. All of these additional devices have important
uses in the control of the centrifugal compressor.
11.1.3.1 Inlet
The inlet to a centrifugal compressor is typically a simple pipe. It may include features such as a valve,
stationary vanes/airfoils (used to help swirl the flow) and both pressure and temperature
instrumentation. All of these additional devices have important uses in the control of the centrifugal
compressor.
11.1.3.2 Centrifugal Impeller
The key component that makes a compressor centrifugal is the centrifugal impeller, Figure 11.1,
which contains a rotating set of vanes (or blades) that gradually raises the energy of the working gas.
This is identical to an axial compressor with the exception that the gases can reach higher velocities
301

and energy levels through the impeller's increasing radius. In many modern high-efficiency
centrifugal compressors the gas exiting the impeller is traveling near the speed of sound. Impellers
are designed in many configurations including "open" (visible blades), "covered or shrouded", "with
splitters" (every other inducer removed) and "w/o splitters" (all full blades). Figure 11.2 show open
impellers with splitters. Most modern high efficiency impellers use "back sweep" in the blade
shape391-392. Euler’s pump and turbine equation plays an important role in understanding impeller
performance.
11.1.3.3 Diffuser
The next key component to the simple centrifugal compressor is the diffuser. Downstream of the
impeller in the flow path, it is the diffuser's responsibility to convert the kinetic energy (high velocity)
of the gas into pressure by gradually slowing (diffusing) the gas velocity. Diffusers can be vaneless,
vane or an alternating combination. High efficiency vane diffusers are also designed over a wide
range of solidities from less than 1 to over 4. Hybrid versions of vane diffusers include: wedge,
channel, and pipe diffusers. There are turbocharger applications that benefit by incorporating no
diffuser. Bernoulli's fluid dynamic principle plays an important role in understanding diffuser
performance.

Figure 11.3 Jet Engine Cutaway Showing the Centrifugal Compressor among others

11.1.3.4 Collector
The collector of a centrifugal compressor can take many shapes and forms. When the diffuser
discharges into a large empty chamber, the collector may be termed a Plenum. When the diffuser
discharges into a device that looks somewhat like a snail shell, bull's horn or a French horn, the

391Japikse, David. “Centrifugal Compressor Design and Performance”. Concepts ETI . ISBN 0-933283-03-2.
392 Aungier, Ronald H. (2000). “Centrifugal Compressors, A Strategy for Aerodynamic Design and Analysis”. ASME
Press. ISBN 0-7918-0093-8.
302

collector is likely to be termed a volute or scroll. As the name implies, a collector’s purpose is to gather
the flow from the diffuser discharge annulus and deliver this flow to a downstream pipe. Either the
collector or the pipe may also contain valves and instrumentation to control the compressor.
11.1.4 Applications
Below, is a partial list of centrifugal compressor applications each with a brief description of some of
the general characteristics possessed by those compressors. To start this list two of the most well-
known centrifugal compressor applications are listed; gas turbines and turbochargers.
11.1.4.1 Gas Turbines and Auxiliary Power Units
In their simple form, modern gas turbines operate on the Brayton cycle. Either or both axial and
centrifugal compressors are used to provide compression. The types of gas turbines that most often
include centrifugal compressors include turboshaft, turboprop, auxiliary power units, and micro-
turbines. The industry standards applied to all of the centrifugal compressors used in aircraft
applications are set by the FAA and the military to maximize both safety and durability under severe
conditions. Centrifugal impellers used in gas turbines are commonly made from titanium alloy
forgings. Their flow-path blades are commonly flank milled or point milled on 5-axis milling
machines. When tolerances and clearances are the tightest, these designs are completed as hot
operational geometry and deflected back into the cold geometry as required for manufacturing. This
need arises from the impeller's deflections experienced from start-up to full speed/full temperature
which can be 100 times larger than the expected hot running clearance of the impeller.
11.1.4.2 Automotive and Diesel Engines Turbochargers and Superchargers
Centrifugal compressors used in conjunction with reciprocating internal combustion engines are
known as turbochargers if driven by the engine’s exhaust gas and turbo-superchargers if
mechanically driven by the engine. Ideal gas properties often work well for the design, test and
analysis of turbocharger centrifugal compressor performance.
11.1.4.3 Natural Gas to Move the Gas from the Production site to the Consumer
Centrifugal compressors for such uses may be one or multi-stage and driven by large gas turbines.
The impellers are often if not always of the covered style which makes them look much like pump
impellers. This type of compressor is also often termed an API-style. The power needed to drive these
compressors is most often in the thousands of horsepower (HP). Use of real gas properties is needed
to properly design, test and analyze the performance of natural gas pipeline centrifugal compressors.
11.1.4.4 Oil Refineries, Natural Gas Processing, Petrochemical and Chemical Plants
Centrifugal compressors for such uses are often one-shaft multi-stage and driven by large steam or
gas turbines. Their casings are often termed horizontally split or barrel. Standards set by the
industry (ANSI/API, ASME) for these compressors result in large thick casings to maximize safety.
The impellers are often if not always of the covered style which makes them look much like pump
impellers. Use of real gas properties is needed to properly design, test and analyze their performance.
11.1.4.5 Air-Conditioning and Refrigeration and HVAC
Centrifugal compressors quite often supply the compression in water chillers cycles. Because of the
wide variety of vapor compression cycles (thermodynamic cycle, thermodynamics) and the wide
variety of workings gases (refrigerants), centrifugal compressors are used in a wide range of sizes
and configurations. Use of real gas properties is needed to properly design, test and analyze the
performance of these machines.
11.1.4.6 Industry and Manufacturing to Supply Compressed Air
Centrifugal compressors for such uses are often multistage and driven by electric motors. Inter-
cooling is often needed between stages to control air temperature. Note that the road repair crew
and the local automobile repair garage find screw compressors better adapt to their needs Ideal gas
303

relationships are often used to properly design, test and analyze the performance of these machines.
Carrier’s equation is often used to deal with humidity.
11.1.4.7 Air Separation Plants to Manufacture Purified End Product Gases
Centrifugal compressors for such uses are often multistage using inter-cooling to control air
temperature. Ideal gas relationships are often used to properly design, test and analyze the
performance of these machines when the working gas is air or nitrogen. Other gases require real gas
properties.
11.1.4.8 Oil Field Re-Injection of High Pressure Natural Gas to Improve Oil Recovery
Centrifugal compressors for such uses are often one-shaft multi-stage and driven by gas turbines.
With discharge pressures approaching 700 bar, casing are of the barrel style. The impellers are often
if not always of the covered style which makes them look much like pump impellers. This type of
compressor is also often termed API-style. Use of real gas properties is needed to properly design,
test and analyze their performance.

11.2 Radial Turbine


A radial turbine is a turbine in which the flow of the working fluid is radial to the shaft393. The
difference between axial and radial turbines consists in the way the fluid flows through the
components (compressor and turbine). Whereas for an axial turbine the rotor is 'impacted' by the
fluid flow, for a radial turbine, the flow is smoothly orientated perpendicular to the rotation axis, and
it drives the turbine in the same way water drives a watermill. The result is less mechanical stress
(and less thermal stress, in case of hot working fluids) which enables a radial turbine to be simpler,
more robust, and more efficient (in a similar power range) when compared to axial turbines. When
it comes to high power ranges (above 5 MW) the radial turbine is no longer competitive (due to heavy
and expensive rotor) and the efficiency becomes similar to that of the axial turbines.

Figure 11.4 Ninety Degree Inward-Flow Radial Turbine Stage

11.2.1 Advantages and Challenges


Compared to an axial flow turbine, a radial turbine can employ a relatively higher pressure ratio
(≈4) per stage with lower flow rates. Thus these machines fall in the lower specific speed and power
ranges. For high temperature applications rotor blade cooling in radial stages is not as easy as in axial
turbine stages. Variable angle nozzle blades can give higher stage efficiencies in a radial turbine stage
even at off-design point operation. In the family of hydro-turbines, Francis turbine is a very well-
known IFR turbine which generates much larger power with a relatively large impeller.

393 From Wikipedia, the free encyclopedia.


304

11.2.2 Types of Radial Turbines


Radial flow turbines may be classified as:
• Inward flow radial (IFR) turbines
➢ Cantilever turbine
➢ 90 degree turbine (see Figure 11.4)
• Outward flow radial (OFR) turbines (see Figure 11.5)
11.2.2.1 Cantilever Radial Turbine
In cantilever IFR turbine the blades are limited to the region of the rotor tip extending from the rotor
in the axial direction. The cantilever blades are usually of the impulse type (or low reaction), such
that there is little change in relative velocity at inlet and outlet of the rotor. Aerodynamically, the
cantilever turbine is similar to an axial-impulse turbine and can even be designed by similar methods.
The fact that the flow is radially inwards hardly alters the design procedure because the blade radius
ratio r2/r3 is close to unity anyway.
11.2.2.2 90 Degree IFR Turbine
The 90° IFR turbine or centripetal turbine is very similar in appearance to the centrifugal
compressor, but with the flow direction and blade motion reversed.

Figure 11.5 Outward Flow Radial Turbine

11.2.2.3 Outward-Flow Radial Stages


In outward flow radial turbine stages, the flow of the gas or steam occurs from smaller to larger
diameters. The stage consists of a pair of fixed and moving blades. The increasing area of cross-
section at larger diameters accommodates the expanding gas. This configuration did not become
popular with the steam and gas turbines. The only one which is employed more commonly is the
Ljungstrom double rotation type turbine. It consists of rings of cantilever blades projecting from two
discs rotating in opposite directions. The relative peripheral velocity of blades in two adjacent rows,
with respect to each other, is high. This gives a higher value of enthalpy drop per stage. (see Figure
11.5).
305
306

12 Best Procedures for Turbomachinery


Here we pay attention to the summary of the most important knowledge and experience a CFD
engineer needs in order to perform CFD simulations of turbo-machinery components394. The guide
is mainly aimed at axial turbo-machinery. The goal is to give a CFD engineer, who has just started
working with turbo-machinery simulations, a head start and avoid some of the most difficult pit-falls.
Experienced turbo-machinery CFD engineers can also use the guidelines in order to learn what other
experts consider best practice. The intended audience is expected to know basic CFD terminology
and have basic turbo-machinery knowledge, but no detailed knowledge about CFD for turbo-
machinery is needed. Before starting a new turbo machinery simulation it is wise to think carefully
of what it is that should be predicted and what physical phenomena that affect the results. This
chapter contains a brief overview of the various types of simulations and some hints of what can be
predicted with them.

12.1 Quasi-3D (Q3D) or 3D Simulation


12.1.1 2D Simulations
These are often used in the early design phase in order to obtain a typical 2D section of a blade. For
cases with many long blades or vanes, like low-pressure turbines, a 2D simulation can also provide
reasonable results. If the area of the flow-path changes significantly in the axial direction it might be
necessary to instead make a quasi-3D simulation.
12.1.2 Quasi-3D (Q3D) Simulation
Two-dimensional flow analyses in the hub-to-shroud and blade-to-blade surfaces to approximate the
3D flow in a blade passage. It is a 2D simulation in which extra source terms are used to account for
the acceleration/deceleration caused by a changing channel height or growing end-wall boundary
layers. Codes focused on turbomachinery applications often have the possibility to perform quasi-3D
simulations, but most general purpose CFD codes cannot do this type of simulations, or require user
coding to implement the correct source terms in the equations.
12.1.3 Full 3D Simulations
Are necessary if a true 3D geometry is needed to obtain correct secondary flows and/or shock
locations. For low-aspect-ratio cases with only a few short blades, like for example structurally
loaded turbine outlet guide vanes, the secondary flow development is important and a 3D simulation
is often necessary in order to obtain reasonable results. For applications where the end-wall
boundary layers grow 3D possibility and you require it. Many codes require special routines or
hidden commands to enable very quickly and interact with a large part of the flow-field it is necessary
to perform a full 3D simulation. This is often the case in compressors and fans, where the negative
pressure gradients make the boundary layers grow much quicker than what they do in for example
turbines. For cases where the shock location is very critical, like in transonic compressors, it is also
often necessary to perform a 3D simulation in order to obtain reasonable shock locations. Figure
12.1 shows the flow range from 2D to 3D by different vendor codes.

394 CFD on line series.


307

3D transient Solver from ANSYS

Q3D Solver from


NUMECA
2-D Steady state
transonic viscous flow

Figure 12.1 Different Flow (2D, Q3D, and full 3D)

12.2 Single vs Multi-Stage Analysis


12.2.1 Single Stage
Many single-stage computations are still performed for turbomachinery design and analysis, and
before the introduction of multi row computations, CFD could only be applied to single blade rows in
isolation. For such computations, it is essential to ensure that the boundary conditions applied are
accurate. These can be extracted from a through-flow computation of the whole machine, and this is
the normal approach for design work, or alternatively, experimental measurements of the inlet and
exit flow field are applied as boundary conditions. The agreement between CFD and experimental
data shown here is better than average. There is a close match in the shape of all the characteristic
curves and the absolute levels of pressure ratio and choking mass flow are accurately reproduced.
However, the stall point is not predicted accurately, and should not be expected to be, since stall is
308

inherently unsteady and involves the full-annulus flow field. Also, at part speeds, the predicted
efficiency values are noticeably lower than the measured values.
12.2.2 Multi-Stage Analysis
12.2.2.1 Steady Mixing-Plane Simulations
Since the mixing-plane method was first introduced in [Denton & Singh 1979] it has become the
industry standard type of rotor-stator simulations. A mixing-plane simulation is steady and only
requires one rotor blade and one stator blade per stage. Between the rotating blade passage and the
steady vane passage the flow properties are circumferentially averaged in a so-called mixing-plane
interface. This will of course remove all transient rotor-stator interactions, but it still gives fairly
representative results. In some commercial codes (CFX for example) mixing-plane interfaces are also
called stage-interfaces.
12.2.2.2 Steady Frozen Rotor Simulations
In a frozen rotor simulation the rotating and the stationary parts have a fixed relative position. A
frame transformation is done to include the rotating effect on the rotating sections. This will give a
steady flow and no transient effects are included. With a frozen-rotor simulation rotating wakes,
secondary flows, leading edge pressure increases etc. will always stay in exactly the same positions.
This makes a frozen rotor simulation very dependent on exactly how the rotors and the stators are
positioned. Most often a mixing-plane simulation gives better results. Frozen rotor simulations are
mainly performed to obtain a good starting flow-field before doing a transient sliding-mesh
simulation.
12.2.2.3 Unsteady Sliding Mesh Stator-Rotor Simulations
This is the most complete type of stator-rotor simulation, and very CPU intensive. In most engines
the number of stators and rotors do not have a common denominator (to avoid instabilities caused
by resonance between different rings). Hence, to make a full unsteady sliding-mesh computation it is
necessary to have a mesh which includes the full wheel with all stators vanes and all rotor blades.
This is often not possible, instead it is necessary to reduce the number of vanes and blades by finding
a denominator that is almost common and then scales the geometry slightly circumferentially. Here
is an example: Real engine: 36 stator vanes, 41 rotor blades Approximated engine: 41 stator vanes,
41 rotor blades, making it possible to simulate only 1 stator vane and 1 rotor blade Scaling of stator:
All stator vanes are scaled by 36/41 = 0.8780 circumferentially.
12.2.2.4 Unsteady Harmonic Balance Simulations
To overcome the computational costs associated with sliding mesh, a technique called Harmonic
Balance is used. The analysis exploits the fact that many unsteady flows of interest in turbomachinery
are periodic in time. Thus, the unsteady flow conservation variables may be represented by a Fourier
series in time with spatially varying coefficients. This assumption leads to a harmonic balance form
of the Euler or Navier–Stokes equations, which, in turn, can be solved efficiently as a steady problem
using conventional computational fluid dynamic (CFD) methods, including pseudo time marching
with local time stepping and multigrid acceleration. Figure 12.2 displays a full analyses blade
solution using a harmonic balanced techniques, courtesy of (CD-Adapco.com). To relax the
fundamental linear assumption while taking advantage of the high solution efficiency, a nonlinear
harmonic method was proposed. Similarly to the time-domain Fourier model, the unsteadiness is
represented by the Fourier series. But now each harmonic will be balanced (‘harmonic balancing’)
respectively in the nonlinear flow equations. Consequently, for a Fourier series retaining N
harmonics, we will have 2N equations for the complex harmonics. In addition, the time-averaged flow
will now be different from the steady flow due to the added deterministic stresses. So in total we have
2N+1 steady-like flow equations, which are solved simultaneously to reflect the interactions between
the unsteady harmonics and the time mean flows. The interactions among the harmonics are
included in a more complete nonlinear harmonic formulation by Hall’s harmonic balance
309

formulations. The nonlinear harmonic approach have been extended to effectively solve rotor-
rotor/stator-stator interactions in multistage turbomachines 395.
12.2.2.5 Hybrid Steady-Unsteady Stator-Rotor Simulations
Hybrid steady-unsteady methods have been proposed in literature in order to have an unsteady
simulation embedded in a multistage steady study. There are several advantages related to this
method: mainly grid size and number of iterations.
12.2.2.6 Other Advanced Multi-Stage Methods
Time-inclined, Adamczyk stresses, etc.

Figure 12.2 Full Blade Simulation using Harmonic Balanced Method (Courtesy of CD-adapco)

12.3 Inviscid or Viscid


For attached flows close to the design point and without any large separations it is often sufficient
with an in-viscid Euler simulation in order to obtain reasonable blade loadings and pressure
distributions. Note that in-viscid Euler simulations should only be used if the boundary layers are
judged to not have a significant effect on the global flow-field. A viscous Navier-Stokes simulation is
necessary in order to predict losses, secondary flows and separations. As soon as separations are of
interest it is of course also necessary to do a viscous simulation. Note that with today’s computers it
is often not time and resources that make users run in-viscid Euler simulations. Running viscous
Navier-Stokes simulations is now so quick that it is not a time problem anymore. Euler simulations
are still interesting though, since with an in-viscid Euler simulation you don't have to worry about
wall resolutions, y+ values, turbulence modeling errors etc.

12.4 Transient or Steady-State


When studying rotor/stator interaction in compressors and turbines there is essentially one choice
that modify dramatically the accuracy of the simulation; whether to perform a steady simulation

395 L. He, T. Chen, R.G.


Wells, Y.S. Li and W. Ning, ‘Analysis of Rotor-Rotor and Stator-Stator Interferences in Multi-
stage Turbomachines’, ASME Journal of Turbomachinery, Vol.124, No.4, pp. 564-571, Oct, 2002.
310

(with mixing plane or similar approaches) or an unsteady one. Although steady simulations with
mixing plane have been extensively performed during the 90s, it must be underlined that the
assumption of a smeared-out field on the rotor/stator interface is too strong for the current request
of accuracy. In fact, different authors396 has shown that the stagnation pressure is representative of
the losses in a steady environment only: in a steady adiabatic case an entropy rise on the streamline
is always associated to a total pressure decrease, while considering an unsteady but inviscid case, the
pressure variations in time influence the stagnation enthalpy. He demonstrated that, in an unsteady
viscous situation, the total pressure variations can provide some information on the global losses but
are also affected by the Euler flow field far from the blade surfaces. Furthermore, Payne et al. 397
individuated a large fluctuation of the time resolved stage efficiency, underlying the importance of
the vane phase on the unsteady losses entity. In addition, [Pullan]398 demonstrated that a steady
simulation generates 10 % less losses compared with the unsteady one.
Another classical error caused by a steady simulation is the analysis of the redistribution of a hot spot
in the rotor row. It was demonstrated that in an axial machine the hot fluid tends to accumulate on
the pressure side of rotor blades. This result can be explained considering that for a steady isentropic
flow without body forces, for a prescribed geometry with a uniform total pressure inlet field, the
streamlines, both the Mach number and the static pressure fields at the vane outlet are not influenced
by the total temperature inlet field. It means that at the stator exit section the hot fluid has a higher
velocity than the surrounding one. Considering the velocity triangles at the rotor inlet, the typical
mechanism of the segregation effect399 is obtained. The so-called “positive jet effect” is an inherently
unsteady phenomenon that interacts with passage vortex: the secondary redistribution brings hot
fluid from suction to pressure side circumferentially across the vane, thus spreading hot fluid over
the entire pressure surface of the blade 400. As a result, the heat load on the blade pressure side is
increased and the life time of the blade reduced by the increased rate of creep. A steady calculation
with mixing plane is not able to reproduce such kind of phenomenon since tangential non-
uniformities at the vane exit section are neglected401. It can be concluded that an accurate unsteady
simulation of the turbine stage should be always done as a support to the steady simulation results.
The unsteady analysis allows to model several important phenomena:

• Unsteady inlet distortions when boundary conditions affect the performances of the gas
turbine;
• Potential interaction caused by the pressure waves travelling (and reflecting) across the
stator/rotor gap;
• Rotating stall: typical of the compressors, is caused by the blockage of some vanes due to the
wrong incidence which causes flow separation;

396 He, L. VKI Lecture Series Part I: Modelling issues for computations of unsteady turbomachinery flows. VKI
Lecture Series on “Unsteady Flows in Turbomachines”, Von Karman Institute for Fluid Dynamics, (1996).
397 Payne, S. J., Ainsworth, R. W., Miller, R. J., Moss, R. W., & Harvey, N. W.,”Unsteady loss in a high pressure

turbine stage: Interaction effects”. International Journal of Heat and Fluid Flow, 26, 695–708, 2005.
398 Pullan, G. (2006). “Secondary flows and loss caused by blade row interaction in a turbine stage”. ASME Journal

of Turbomachinery, 128(3), 484–491.


399 Kerrebrock, J. L., & Mikolajczak, A. A. (1970). “Intra-stator transport of rotor wakes and its effect on

compressor performance”. ASME Journal of Engineering for Power, 92(4), 359–368.


400 Dorney, D. J., Davis, R. L., Edwards, D. E., & Madavan, N. K. (1992). “Unsteady analysis of hot streak migration

in a turbine stage”. AIAA Journal of Propulsion and Power, 8(2), 520–529.


401 F. Montomoli et al., “Uncertainty Quantification in Computational Fluid Dynamics and Aircraft Engines”,

SpringerBriefs in Applied Sciences and Technology, DOI 10.1007/978-3-319-14681-2_2. Chapter 2 -


Limitations in Turbomachinery CFD , 2015.
311

• Wake passing: is fundamental in low-pressure turbines for the suppression of laminar


separation bubbles;
• Aero-elastic instability: generally called “flutter”, is generated by the blade mechanical
response to the unsteady disturbances.

All said, because of inherent difficulty, today most turbo-machinery simulations are performed as
steady-state simulations. Transient simulations are done when some kind of transient flow behavior
has a strong influence on the global flow field. Examples of transient simulations are detailed
simulations of rotor-stator interaction effects, simulations of large unsteady separations etc.
Sometimes when you perform
a steady stationary simulation
you can see tendencies of
unsteady behavior like for
example periodic vortex
shedding behind blunt trailing
edges. This is often first seen
as periodical variations of the
residuals. If the unsteady
tendencies are judged to not
affect the overall simulation
results it might be necessary
to coarsen the mesh close to
the vortex shedding or run a
different turbulence model in
order to make the simulation
converge. Sometimes you are
still forced to run a transient Figure 12.3 Transient Blade Row Extensions Enable Efficient Multi-
simulation and average the Stage CFD Simulation (Courtesy of ANSYS.com)
results if you don't obtain a
converged steady solution. Figure 12.3 shows the transient blade row extensions enable efficient
multi-stage CFD simulation (courtesy’s of ANAYS.com).

12.5 Meshing
In turbomachinery applications structured multi-block hexahedral meshes are most often used for
flow-path simulations. In most solvers a structured grid requires less memory, provides superior
accuracy and allows a better boundary-layer resolution than an unstructured grid. By having cells
with a large aspect ratio around sharp leading and trailing edges a structured grid also provides a
better resolution of these areas. Many companies have automatic meshing tools that automatically
mesh blade sections with a structured mesh without much user intervention. Unstructured meshes
are used for more complex and odd geometries where a structured mesh is difficult to create. Typical
examples where unstructured meshes are often used are blade tip regions, areas involving leakage
flows and secondary air systems, film cooling ducts etc. When meshing avoid to create large jumps
in cell sizes. Typically the cell size should not change with more than a factor of 1.25 between
neighboring cells. For structured meshes also try to create fairly continues mesh lines and avoid
discontinuities where the cell directions suddenly change. For multi-block structured meshes avoid
placing the singular points where blocks meet in regions with strong flow gradients since most
schemes have a lower accuracy in these singular points. Figure 12.4 shows a typical meshing for a
312

turbomachinery stage. dynamics. It must


be underlined that a “perfect mesh” does
not exist. Once the outcome of the
numerical activity has been decided, a
proper definition of the mesh parameters
to capture the essential flow properties
must be devised. Furthermore, mesh
quality must be coherent with the selected
numerical approach with special
attention to steady/unsteady analysis and
to turbulence modelling402.
12.5.1 Mesh Size Guidelines
It is difficult to define, a priori, the mesh
size. The required mesh size depends on
the purpose of the simulation. If the main
goal is to obtain static pressure forces a Figure 12.4 Typical Meshing of a Turbomachinery
coarse mesh is often able to obtain a good Stage
solution, especially when an accurate resolution of the boundary layers is not required. For 2D in-
viscid simulations of one blade a mesh with say 3,000 cells is most often sufficient. For 3D in-viscid
blade simulations a mesh size of about 40,000 cells is usually sufficient. On in-viscid Euler simulations
the cells should be fairly equal in size and no boundary layer resolution should be present. Avoid
having too skewed cells. For loss predictions and cases where boundary layer development and
separation is important the mesh needs to have a boundary layer resolution. The boundary layer
resolution can either be coarse and suitable for a wall function simulation or very fine and suitable
for a low-Re simulation. For further information about selecting the near-wall turbulence model
please see the turbulence
modeling section. In 3D single-
blade simulations a decent
wall-function mesh typically
has around 100,000 cells. This
type of mesh size is suitable for
quick design iterations where it
is not essential to resolve all
secondary flows and vortices.
A good 3D wall-function mesh
of a blade section intended to
resolve secondary flows well
should have at least 400,000
cells. A good low-Re mesh with
resolved boundary layers
typically has around 1,000,000
cells. In 2D blade simulations a
Figure 12.5 Multi-Block Grid for the Space Shuttle Main Engine
good wall-function mesh has
Fuel Turbine
around 20,000 cells and a good
low-Re mesh with resolved
boundary layers has around 50,000 cells. Along the suction and pressure surfaces it is a good use

402F. Montomoli et al., “Uncertainty Quantification in Computational Fluid Dynamics and Aircraft Engines”,
SpringerBriefs in Applied Sciences and Technology, DOI 10.1007/978-3-319-14681-2_2. Chapter 2-
Limitations in Turbomachinery CFD , 2015.
313

about 100 cells in the stream wise direction. In the radial direction a good first approach is to use
something like 30 cells for a wall-function mesh and 100 cells for a low-Re mesh. It is important to
resolve leading and trailing edges well. Typically at least 10 cells, preferably 20 should be used
around the leading and trailing edges. For very blunt and large leading edges, like those commonly
found on HP turbine blades, 30 or more cells can be necessary. Cases which are difficult to converge
with a steady simulation and which show tendencies of periodic vortex shedding from the trailing
edge, can sometimes be "tamed" by using a coarse mesh around the trailing edge. This, of course,
reduces the accuracy but can be a trick to obtain a converged solution if time and computer resources
do not allow a transient simulation to be performed. Figure 12.5 shows a multi-block grid for the
space shuttle main engine fuel turbine (AIAA 98-0968). As an example, several schemes for steady
analysis of a two-dimensional profile could lead to a non-converged solution when the spatial
resolution in the trailing edge region is too fine, since an unstable base region could occur despite the
steady assumption. At the same time, it is wrong to perform a large-eddy simulation with a coarse
mesh, since the sub-grid scale model would try to account for the vortex structures. It is worth
mentioning also the evaluation of the boundary layer development, which is strongly dependent on
both the selected model and the near wall mesh resolution403.
12.5.2 Case Study - Mesh Resolution Effect on 3D RANS Turbomachinery Flow Simulations
Over the past twenty years, guidelines for choosing the mesh resolution for the numerical simulation
of turbomachinery viscous flows using the RANS models have changed several times: from 100-200
K cells per one blade in 90s to 0.5 -1.0 M cells per one blade now. Usually, a mathematical basis of
such recommendations is not clear, requirements to the mesh refinement are often not well-founded
(perhaps, with the only exception for y+) and the question of the solution convergence remains open.
Recently, a new investigation by [Yershov & Yakovlev]404 presented the study of the effect of a mesh
refinement on numerical results of 3D RANS computations of turbomachinery flows. The CFD solver
based on the second-order accurate scheme. The simplified multigrid algorithm and local time
stepping permit decreasing computational time. The flow computations are performed for a number
of turbine and compressor cascades and stages. In all flow cases, the successively refined meshes of
H-type with an approximate orthogonality near the solid walls were generated. The results obtained
are compared in order to estimate their both mesh convergence and ability to resolve the transonic
flow pattern. It is concluded that for thorough studying the fine phenomena of the 3D
turbomachinery flows, it makes sense to use the computational meshes with the number of cells from
several millions up to several hundred millions per a single turbomachinery blade channel, while for
industrial computations, a mesh of about or less than one million cells per a single turbomachinery
blade channel could be sufficient under certain conditions.
12.5.2.1 Formulation of Problems
We have performed the RANS simulation of the 3D turbulent compressible viscous flow through
several turbomachinery stages and cascades where the k–ω SST turbulence model was used. The
main objective was a qualitative study of the numerical solution convergence without being tied to
the experimental data. It is evident that both the insufficient adequacy of the mathematical model as
well as approximation and experimental errors could lead to the fact that in some cases the
differences between the numerical results and the experimental data may increase as the mesh is
refined. The meshes considered in this study were conventionally divided into five groups based on
the number of cells per one blade channel:

403 F. Montomoli et al., “Uncertainty Quantification in Computational Fluid Dynamics and Aircraft Engines”,
SpringerBriefs in Applied Sciences and Technology, DOI 10.1007/978-3-319-14681-2_2. Chapter 2-
Limitations in Turbomachinery CFD , 2015.
404 Sergiy Yershov, Viktor Yakovlev, “Mesh Resolution Effect On 3d Rans Turbomachinery Flow Simulations”,

Institute For Mechanical Engineering Problems Of Nasu, Kharkiv, Ukraine, 2016.


314

1- very coarse meshes of 104–105 cells;


2- coarse meshes of 105–106 cells;
3- intermediate meshes of 106–107 cells;
4- fine meshes of 107–108 cells;
5- very fine meshes of 108–109 cells.
During the present mesh convergence study the number of cells in each spatial direction increased
about twice. For the meshes of all considered groups value of y+ both in the radial and circumferential
directions was set close to unity. It was found that an adequate prediction of the law of the wall
(universal velocity profile) is possible only if at least 30 cells are placed across the boundary layer
and the mesh expansion ratio in the wall-normal direction does not exceed 1.1. Therefore, when we
performed computations on meshes of the groups 3, 4, and 5, these requirements were strictly
enforced. Ensuring these requirements on meshes of the groups 1 and 2 without reducing the
accuracy in the flow core is very problematic at best. It should be noted that the computations using
meshes of more than one million cells require considerable computational time and are almost
impossible without the mechanisms of the convergence acceleration implemented in the solver.
12.5.2.2 Conclusions
The present study confirms the well-known fact that the mesh scales should match the flow scales,
namely the characteristic size of the flow regions with significant gradients of thermodynamic,
kinematic and turbulent parameters. On the other hand, the obtained results show that the mesh
convergence of the kinetic energy losses requires sufficiently fine meshes of 107 cells and more per
one blade channel when using the second order-accurate numerical scheme. A good resolution of
shock waves, separation zones, wakes, and tangential discontinuities needs the same meshes. An
additional mesh refinement may be necessary due to various small-scale features of flow or flow path
geometry, such as film cooling holes, vortex generators, etc. It should be emphasized again that all
aforesaid concerns to the numerical solutions of the RANS equations and more sophisticated
turbulent flow models, such as DNS and LES, require further special researches. For detail
explanation of meshing strategy, readers should review [Yershov & Yakovlev]405.
Scientific researches of the fine flow patterns require a high accuracy and a detail resolution, so, in
this case, preference should be given to fine or very fine meshes of the groups 4 and 5. Since such
computations is very time consuming, it may be considered acceptable to use intermediate meshes
of the group 3, if the mentioned above requirement on the near-wall cell size, the number of cells
across boundary layer, and the mesh expansion ratio in the wall-normal direction are satisfied. In the
case of high-volume industrial computations, the use of intermediate (of the group 3) or even coarser
meshes sometimes may be sufficient. However, it should be remembered that such computations
often result in the flow path efficiency increase by only 0.001-0.002 (0.1-0.2 percentage points),
which is comparable with or even less than discretization errors. Therefore, the final results of such
computations should be always verified using finer meshes.
12.5.3 Boundary Mesh Resolution
For design iteration type of simulations where a wall function approach is sufficient y+ for the first
cell should be somewhere between 20 and 200. The outer limit is dependent on the actual Re number
of the simulations. For cases with fairly low Re numbers make sure to keep the maximum y+ as low
as possible. For more accurate simulations with resolved boundary layers the mesh should have a y+
for the first cell which is below 1. Some new codes are now using a hybrid wall treatment that allows
a smooth transition from a coarse wall-function mesh to a resolved low-Re mesh. Use some extra care
when using this type of hybrid technique since it is still fairly new and unproven. Outside of the first
cell at a wall a good rule of thumb is to use a growth ratio normal to the wall in the boundary layer of

405Sergiy Yershov, Viktor Yakovlev, “Mesh Resolution Effect On 3d Rans Turbomachinery Flow Simulations”,
Institute For Mechanical Engineering Problems Of Nasu, Kharkiv, Ukraine, 2016.
315

maximum 1.24. For a low-Re mesh this usually gives around 40 cells in the boundary layer whereas
a wall-function mesh does not require more than 10 cells in the boundary layer. If you are uncertain
of which wall distance to mesh with you can use a y+ estimation tool to estimate the distance needed
to obtain the desired y+. These estimation tools are very handy if you have not done any previous
similar simulations. As a rule of thumb a wall-function mesh typically requires around 5 to 10 cells
in the boundary layer whereas a resolved low-Re mesh requires about 40 cells in the boundary layer.
12.5.4 Periodic Meshing
To reduce the time, efforts and complexity of meshing the rotational periodicity of the impeller
geometry is taken advantage. Axial machines and rotating fluid zone of radial & mixed flow machines
are meshed using this approach. Choosing a single periodic flow passage is the first step in this
approach. The periodic angle of the flow passage is decided by the number of vanes/blades present.
For example, Periodic angle or Angle of Rotational Periodicity = 360°/number of blades.
• For a radial turbine with 16 blades, Angle of rotational periodicity → 360°/16 =22.5° (single
blade passage)
• For a pump with 4 blades, Angle of rotational periodicity → 360°/4 = 90° (single blade
passage)
This periodic geometric sector can be chosen in two different ways.
• Flow passage between two blades (suction side of first blade to the pressure side of next
blade).
• To have one complete blade inside the periodic flow passage.
There are two different scenarios based on the flow physics. If the flow physics is also periodic (most
axial flow machines), the mesh is generated only for a single blade fluid passage (ϴ), regardless of the
number of blades and is directly used for simulation. But if the flow physics is not periodic (radial &
mixed flow machines with volute), the mesh is generated for the single periodic flow passage / sector
and is copy rotated to get mesh for the complete geometry (360°). Meshing software provides an
option for periodic meshing to ensure both sides of periodic passage has same number of nodes and
same node location with a rotational offset of ϴ.

12.6 Boundary Conditions


Turbomachinery CFD employs multi-region approach, the computational model is split into a
number of regions. Any number of regions is allowed. Each region has its own independent mesh
Turbomachinery CFD employs multi-region approach, the computational model is split into a
number of regions. Any number of regions is allowed. Each region has its own independent mesh
and case set-up. Regions are like serial connected and communicate via interfaces. Typically, velocity
is prescribed at the inlet and pressure is prescribed at the outlet. Describing different types of
boundary conditions and when they should be used not easy as it sound. For each of the analysis
methods, boundary conditions must be specified at the inlet and exit of the computational domain.
In addition, for averaging plane methods, average flow properties must be transferred between the
blade rows at grid interfaces406. It is common practice to force the flow to be axisymmetric at these
boundaries. Although axisymmetric boundary conditions are simple to apply and tend to be
numerically robust, they can reflect outgoing waves and thereby hinder convergence and
contaminate the interior solution. Axisymmetric boundary conditions can be particularly bad at the
inlet of transonic compressors or at the exit of transonic turbines, and between closely-spaced blade

406Rodrick V. Chima, “Calculation of Multistage Turbomachinery Using Steady Characteristic Boundary


Conditions”, AIAA 98-0968.
316

rows. [Giles]407 presented a unified theory for the construction of Non-reflecting boundary conditions
for the Euler equations (NRBC’s). The boundary conditions are based on the linearized Euler
equations written in terms of perturbations of primitive variables about some mean flow. Wave-like
solutions are substituted into the flow equations, and the solution is circumferentially decomposed
into Fourier modes. The zeroth mode corresponds to the mean flow and is treated according to one-
dimensional characteristic theory. This allows average changes in incoming characteristic variables
to be specified at the boundaries. Simply put, since the numerical solution is calculated on a truncated
finite domain, and one must prevent any nonphysical reflections of outgoing waves at the far-field
boundaries that could contaminate the numerical solution.
This becomes essential in turbomachinery applications in which the boundaries are often not very
far from the blades, because the physical spacing between the blade rows can be quite small. It
therefore becomes highly important for an accurate simulation to construct nonreflecting boundary
conditions (NRBCs). Preventing
spurious reflections that would
corrupt the solution is not only
important to get an accurate
prediction of the flow field, but
also to get more efficient
computations; convergence rate
is enhanced due to an
improvement of the transmission
of outgoing waves, allowing
smaller meshes to be used408.
Figure 12.6 compares a contour
plot the pressure Vs. y-
coordinate for both Riemann BC
and NRBC’s in a supersonic
cascade for a 2-D flow [F. De
Raedt, 2015]. The most notable
observation is that at the outflow,
when Riemann BC are applied,
the pressure lines diverge from
the boundary and rarely cross
that boundary. This is a direct Figure 12.6 Pressure contour plot, 2nd order spatial
result of the reflectivity of the discretization scheme
boundary conditions. When the
boundary is far away from the airfoil, the effectiveness of these reflections on the airfoil flow-field is
minimal, as observed by comparing the long flow-field simulations of the Riemann BC and NRBC.
However when the boundary is close to the airfoil, the simulations using Riemann BC become
completely inaccurate. In contrast the short flow-field simulations using the NRBC result in very
similar pressure contours to those of the long flow-field. This clearly demonstrates the effect of the
NRBC implementation. One can have a closer look at the boundary itself to further clarify this
comparison409.

407 Giles, Michael B., “Nonreflecting Boundary Conditions for Euler Equation Calculations,” AIAA Journal, Vol. 28,
No. 12, Dec. 1990, pp. 2050-2058.
408 “Three-Dimensional Nonreflecting Boundary Conditions for Swirling Flow in Turbomachinery”, Journal of

Propulsion and Power Vol. 23, No. 5, September–October 2007.


409 F. De Raedt, “Non-Reflecting Boundary Conditions for non-ideal compressible fluid flows”, Master of Science at

the Delft University of Technology, 2015.


317

The exact knowledge of boundary conditions for numerical simulations is probably one of the most
challenging problems in CFD and it is crucial in turbomachinery. turbomachinery components are
subjected to non-uniform conditions whose distributions have to be determined with high accuracy.
A typical example of this kind of problems is the simulation of a high-pressure stage with realistic
inlet conditions. Salvadori et al410 demonstrated that a non-uniform inlet temperature profile,
including hot streak migration, generates 10% variation in blade suction side static pressure
distribution at mid-span, a 60 % variation of Nusselt number value on blade pressure side and a 19
% variation in the peak total temperature at mid-span at the stage exit section with respect to cases
with uniform inlet. Considering that the distribution of turbine entry temperature is not measured
directly and that an error of more than 50 K is common in real gas turbines, it is clear the impact of
such parameter411.

12.7 Turbulence Modeling


Selecting a suitable turbulence model for turbo-machinery simulations can be a challenging task.
There is no single model which is suitable for all types of simulations. Which turbulence model CFD
engineers use has as much to do with beliefs and traditions as with knowledge and facts? There are
many different schools. However, below follows some advices that most CFD engineers in the turbo-
machinery field tend to agree upon. For attached flows close to the design point a simple algebraic
model like the Baldwin-Lomax model can be used. Another common choice for design-iteration type
of simulations is the one-equation model by Spalart-Allmaras. This model has become more popular
in the last years due to the many inherent problems in more refined two-equation models. The big
advantage with both the Baldwin-Lomax model and the Spalart-Allmaras model over more advanced
models is that they are very robust to use and rarely produce completely unphysical results. In order
to accurately predict more difficult cases, like separating flows, rotating flows, flows strongly affected
by secondary flows etc. it is often necessary to use a more refined turbulence model. Common choices
are two-equation models like the k-ε model. Two-equation models are based on the Boussinesq eddy
viscosity assumption and this often leads to an over-production of turbulent energy in regions with
strong acceleration or deceleration, like in the leading edge region, regions around shocks and in the
suction peak on the suction side of a blade. To reduce this problems it is common to use a special
model variant using, for example, Durbin's realizability constraint or the Kato-Launder modification.
Note that different two-equation models behave differently in these problematic stagnation and
acceleration regions. Worst is probably the standard κ-ε, model, κ-ω model are slightly better but still
do not behave well. More modern variants like Menter's SST κ-ω model also has problems, whereas
the v2f model by Durbin behaves better.

12.8 Aero-Mechanics
Now let’s look at the challenges of aeromechanics. Whereas the aerodynamicist generally prefers
designs with very thin blades, the structural engineer prefers thick blades to minimize stress and
optimize vibration characteristics. Those interested in material cost and weight would no doubt side
with the aerodynamicist, whereas those responsible for honoring the machine warranty would favor
the structural viewpoint. Achieving agreement requires a balance, and that is where the field of
aeromechanics comes in. Aeromechanics is by no means new. What is new is the fidelity with which
engineers can practically consider both the fluid mechanics and the structural aspects of the solution.
The real behavior of rotating blades is indeed very complex, and the mechanical loads are very high.

410 Salvadori, S., Montomoli, F., Martelli, F., Chana, K. S., Qureshi, I., & Povey, T. “Analysis on the effect of a
nonuniform inlet profile on heat transfer and fluid flow in turbine stages”. Journal of Turbomachinery, 134(1),
011012-1-14. doi:10.1115/1.4003233, (2012).
411 F. Montomoli et al., “Uncertainty Quantification in Computational Fluid Dynamics and Aircraft Engines”,

SpringerBriefs in Applied Sciences and Technology, DOI 10.1007/978-3-319-14681-2_2, 2015.


318

For example, a single low-pressure steam turbine blade rotating at operating speed generates a load
of several hundred tons! Long, thin blades are susceptible to vibration. Engineers strive to design
blades whose natural frequencies do not coincide with the disturbances that arise due to operating
speed, etc. That is complicated enough, but there are also periodic disturbances that can originate
from more distant blade rows or aerodynamic effects412.
In the past, analysis of fluids and structural dynamics was mostly separate and simplified. But for
some time, at least in principle, the ability to perform high-fidelity coupled analysis has been
available. In reality, solving for time-dependent, three-dimensional fluid-structure interaction is very
time-consuming and expensive, even on today’s high performance computing systems. Engineers
have opted for more practical, usually disconnected and often lower-fidelity analysis methods.
Recently, practical yet high-fidelity multiple physics solution methods have emerged.

Figure 12.7 Analysis provided vibration required for flutter analysis

Prediction of aerodynamic blade damping, or “flutter,” is one such method. The procedure is to first
solve for the mechanical modes of vibration, and then feed that information to the CFD simulation.
The unsteady CFD simulation deforms the blade in the presence of the flow field and predicts
whether the blade is aerodynamically damped, and hence stable, or not. This high-fidelity approach
is practical because it provides a solution to the full wheel (all of the many blades in a given row) by
solving only for one or at most a few blades in the blade row of interest. Cyclic symmetry is the
enabling structural technology here, while the Fourier Transformation method is key on the CFD side.
Tightly coupled these two efficient methods provides great advances in computing fidelity and speed.
Predicting forced response is essentially the inverse workflow to flutter. Figure 12.7 shows where
analysis provides the mechanical modes of blade vibration required for flutter analysis. First, the
unsteady fluid dynamic loads are predicted, and made available to the structural solver. After a
mechanical harmonic response simulation, the engineer evaluates the results for acceptable levels of
blade displacement, strain and stress. The concept of Nodal Diameter is explained next.

412 ANSYS Blog.


319

12.8.1 Nodal Diameter


Natural frequency is the frequency at which an object vibrates when excited by force. At this
frequency, the structure offers the least resistance to a force and if left uncontrolled, failure can occur.
Mode shape is deflection of object at a given natural frequency. A guitar string is a good example of
natural frequency and mode shapes. When struck, the string vibrates at a certain frequency and
attains a deflected shape. The eigenvalue
(natural frequency) and the accompanying
eigenvector (mode shape) are calculated to
define the dynamics of a structure. A turbine
bladed disk has many natural frequencies and
associated mode shapes. In the case of a bladed
disk, the mode shapes have been described as
nodal diameters. The term nodal diameter is
derived from the appearance of a circular Figure 12.8 Examples of Nodal Diameter
geometry, like a disk, vibrating in a certain mode.
Mode shapes contain lines of zero out-of-plane displacement which cross the entire disk as shown in
Figure 12.8. In other words, a node line is a line of zero displacement and the displacement is out
of phase on the sides of the line represented by white and gray shades in Figure 12.8. These are
commonly called nodal diameters. Hence the natural frequency and nodal diameter are required to
describe a bladed disk mode413.

12.9 Near Wall Treatment


For on-design simulations without any large separated regions it is often sufficient to use a wall-
function model close to the wall, preferably with some form of non-equilibrium wall-function that is
sensitized to stream wise pressure gradients. For off-design simulation, or simulations involving
complex secondary flows and separations, it is often necessary to use a low-Re model. There exist
many low-Re models that have been used with success in turbo machinery simulations. A robust and
often good choice is to use a one-equation model, like for example the Wolfstein model, in the inner
parts of the boundary layer. There are also several Low-Re κ-ε models that work well. Just make sure
they don't suffer from the problem with overproduction of turbulent energy in regions with strong
acceleration or deceleration. In the last few years Menter's low-Re SST κ-ω model has gained
increased popularity.

12.10 Transition Prediction


Transition refers to the process when a laminar boundary layer becomes unstable and transitions to
a turbulent boundary layer. There are two types of transition - natural transition, where inherent
instabilities in the boundary layer cause the transition and by-pass transition, where convection and
diffusion of turbulence from the free-stream into the boundary layer cause the transition. Most
transitions in turbo machinery are by-pass transitions caused by free-stream turbulence and other
external disturbances like wakes, vortices and surface defects. Simulating transition in a CFD code
accurately is very difficult. Often a separate transition model needs to be solved in order to specify
the transition location and length. Predicting natural transition in a pure CFD code is not possible.
Predicting by-pass transition in a pure CFD code is almost impossible, although there are people who
claim to be able to predict by-pass transition with low-Re two-equation models. However, this is
usually on special test cases and with simulations that have been tuned for these special cases, see
for example [Saville 2002]. In reality transition is a very complex and sensitive process where

413Mohamed Hassan, “Vibratory Analysis of Turbomachinery Blades”, Master of Mechanical Engineering,


Rensselaer Polytechnic Institute, Hartford, Connecticut, December, 2008.
320

disturbances like incoming wakes and vortices from previous stages, surface roughness effects and
small steps or gaps in the surfaces play a significant role.
The turbomachinery codes that have transition prediction models often use old ad-hoc 414 models like
the Abu-Ghannam and Shaw model or the Mayle model. These models can be quite reliable if they
have been validated and tuned for a similar application. Do not trust your transition predictions
without having some form of experimental validation. Menter has also recently developed a new form
of transition model that might work fairly well, but it is still too new and untested. For some turbo
machinery applications, like modern high-lift low-pressure turbines, transition is critical. For these
applications a CFD code with a transition model that has been tuned for this type of applications
should be used.

12.11 Numerical Consideration


Use at least a second order accurate scheme for the flow variables. Some codes require a first order
scheme for the turbulent variables (κ-ε) in order to converge well. It might be sufficient with a first
order scheme only on the turbulence variables, but a second order scheme is of course preferable.

12.12 Convergence Criteria


To know when a solution is converged is not always that easy. You need some prior experience of
your CFD code and your application to judge when a simulation is converged. For normal pure aero
simulations without resolved walls, i.e. with wall functions or in-viscid Euler simulations,
convergence can most often be estimated just by looking at the residuals. Exactly what the residuals
should be is not possible to say, it all depends on how your particular code computes and scales the
residuals. Hence, make sure to read the manuals and plot the convergence of a few global parameters
before you decide what the residuals should be for a solution to converge. Note also that many
manuals for general purpose CFD codes list overly aggressive convergence criteria that often produce
un-converged results. For simulations with resolved walls it is good to look at the convergence of
some global properties, like total pressure losses from the inlet to the outlet. For heat transfer
simulations it is even trickier since the aerodynamic field can look almost converged although the
thermal field is not converged at all. If doing heat transfer simulations make sure to plot the heat-
transfer, run for some time, and plot it again to make sure that it doesn't change anymore. With very
well resolved walls and heat transfer it can sometimes take 10 times longer for the thermal field to
converge.

12.13 Single or Double Precision


With today’s computers and cheap memory prices it often does not cost much extra to run in double
precision. Before using single precision you should first investigate how your software and hardware
works with double precision. If the extra time and memory needed for double precision is negligible
you should of course always run in double precision. With double precision you never have to worry
about round-off errors. Always using double precision is one way of avoiding one type of pit-falls in
the complex world of CFD simulations. Use double precision when you have resolved boundary layers
(Y+ around 1) and when you use advanced physical models like combustion, free-surface simulations,
spray and transient simulations with quick mesh motions.

12.14 Heat Transfer Prediction


Besides listing the general heat transfer mechanisms involved (namely conduction, convection,
radiation), heat transfer prediction in CFD may be seen as or split into two cases. Mesh consists of
fluid domain(s): you want to know how heat inserted into the fluid domain (e.g. via the flowing

414 Latin phrase meaning “for this purpose”.


321

medium) changes the temperature field along the flow path. Be it a steady or transient run, a fluid
which enters with a given temperature will usually experience temperature variations, for instance,
caused by convective boundary conditions, prescribed temperature profiles at flow obstacles etc.
Mesh consists of fluid and solid domain(s): additionally to the above, you want info too w.r.t .the
(spatial, temporal or even spectral) temperature field distribution in surrounding, confining or
immersed solids like channel walls or heat exchanger tubes. This is also called conjugate heat transfer
CHT in the CFD context. CHT requires a good boundary layer resolution; usually the wall mesh needs
to be rather refined, to obtain realistic heat flux results at the fluid/solid interface. Flow and heat
transfer convergence require different time step settings, to properly capture changes in flow and
heat quantities respectively. In either case, verify (strict necessity depends on CFD code used, CFX
for instance checks and assists in regard) that model dimensions, boundary conditions and
properties are in consistent units, hold appropriate values. Check temperature-dependence of
properties and other numbers before the run. In heat transfer predictions (depending on the CFD
code in use) besides the flow solver, you may have to activate the thermal solver too, as a job
specification.
12.14.1 Keeping it Cool in Gas Turbine
One of the key problems in the design of advanced gas turbine engines is the development of effective
cooling methods for the turbine vanes and blades415. Due to competition and continuous
improvement an increased complexity of cooling technology is required in the design of turbine
engine parts. In view of the material and time costs for experimental research, CFD has been accepted
by turbomachinery companies as one of the main methods for evaluating the performance of new
designs. Industrial CFD applications range from classical single and multi-blade-row simulations in
steady and transient mode to heat transfer and combustion chamber simulations. Depending on the
type of machine, physical and geometrical effects have to be taken into account. A complicating factor
is that it is necessary to carry out parametric studies considering several geometric options in the
process of designing the cooling systems. This normally takes a lot of time to generate mesh models
due to the mesh resolution required in the boundary layer.

12.15 Literature Review and Parallel Processing Tools


A review of CFD analysis for turbo-machineries is acquired by416. The main points and contributions
are listed below:
➢ Provided a critical review of CFD analysis for turbines, compressors and centrifugal pumps.
➢ Various issues related to the CFD software used in turbomachinery are identified.
➢ Parallel computing tools adopted for parallelization of CFD software used in turbomachinery
are earmarked.
Form the literature survey it is found that, the future of turbomachinery designs will depend even
more extremely on CFD, than they do currently, as the ability of CFD to predict the behavior of fluid
flow and heat transfer is continuously improving. The conclusions drawn from the survey are as
follows:
• At present, the trend is to move from single stage and steady flow to multistage and unsteady
predictions, both of which demands intense computational power.

415 A.V. Rubekina, A.V. Ivanov, G.E. Dumnov, A.A. Sobachkin (Mentor Graphics Corp., Russia); K.V. Otryahina
(PAO NPO Saturn, Russia), “Keeping it Cool in Gas Turbines”, Mentor paper.
416 Runa Nivea Pinto , Asif Afzal , Loyan Vinson D'Souza , Zahid Ansari , Mohammed Samee A. D., “Computational

Fluid Dynamics in Turbomachinery: A Review of State of the Art”, Archives of Computational Methods in
Engineering , DOI: 10.1007/s11831-016-9175-2, April 2016.
322

• The results obtained from unsteady flow can be employed to generate the unsteady blade
loading which indeed helps in prediction of mechanical issues of blading.
• The unsteady flow predicts the loss generation which is currently in use and need to be
explored further.
• With CFD, the prediction of generation of entropy when vortices and wakes mix in unsteady
flow has become definitely very easy than the experiment; but for validation of CFD results,
high class experimental results are required and to interpret the results from CFD.
• Numerical and modeling errors need to be reduced further.
• The accuracy of CFD results obtained for turbulence flow need to be increased.
• The assumptions and approximations made in CFD analysis of turbomachinery has to be
developed for every flow and geometric conditions.

Another important aspect understood from this survey is that the unsteady flow computations need
huge time to obtain results417. This is due to the enormous computational time required by the CFD
software to generate converged results. In order to reduce the computational time of the codes are
parallelized to run on various multi-core/multi-processor architectures. The major parallel
computing tool popularly used for the parallelization of CFD codes in turbomachinery is MPI. Various
parallel computing tools that can be employed further to attempt parallelization of these codes are
provided below:

1. OpenMP (Open Multiprocessing)


2. CUDA (Compute Uni_ed Device Architecture)
3. PVM (Parallel Virtual Machine)
4. Hybrid OpenMP+MPI
5. Tri-level hybrid OpenMP+MPI+CUDA
6. OpenCL (Open Computing Language)

So far, very less effort have been reported in literature regarding parallelization of turbomachinery
related CFD codes employing parallel computing tools mentioned above. These tools have been
proved to be of great computational performance enhancers for the CFD software used in other fields.
More specifically Open MP+MPI hybrid parallel computing approach has shown highly improved
computational performance.

12.16 Concluding Remarks


In previous sections, the limitations introduced by CFD in the analysis of turbomachinery
components have been shown with the most promising approaches used to overcome those
problems. Whatever is the selected approach, there will always be the limitation connected with the
numerical model used to perform the simulation. The accuracy of a numerical simulation is a
combination of the order of accuracy of the discrete equation, of the selected discretization method
(forward/central/backward) and of the order of reconstruction of the gradients. Furthermore, there
will be effects related to the computational mesh (spatial filter) and to the selected time step (time
filter), not to mention the Courant number for dual time stepping approaches. The latter parameter
will also play a role in the selection of the explicit/implicit algorithm, which is also connected with
the accuracy of the model in resolving turbulence. In fact, turbulence is the key problem in
turbomachinery flows since it is possible to range from algebraic methods to direct numerical

417F. Montomoli et al., “Uncertainty Quantification in Computational Fluid Dynamics and Aircraft Engines”,
SpringerBriefs in Applied Sciences and Technology, DOI 10.1007/978-3-319-14681-2_2. Chapter 2-
Limitations in Turbomachinery CFD , 2015.
323

simulation with increasing accuracy and computational costs. It can be also underlined that in
presence of multiphase flows, i.e. for cavitation, and combustion, the selected methodology will
introduce a specific limitation on the obtained result whose entity is hard to be quantified. Most of
the cited problems are related to the stochastic uncertainty, also called reducible uncertainty because
it is a function of the knowledge of the problem physics and of the complexity of the algorithm. Then,
numerical accuracy can rise with an improved knowledge of the physics and with the computational
resources. However, user’s knowledge represents the most important assurance for a good CFD,
while uncertainty quantification is a strong support in the analysis and design of turbomachinery418.

418F. Montomoli et al., “Uncertainty Quantification in Computational Fluid Dynamics and Aircraft Engines”,
Springer Briefs in Applied Sciences and Technology, DOI 10.1007/978-3-319-14681-2_2. Chapter 2-
Limitations in Turbomachinery CFD , 2015.

You might also like