You are on page 1of 221

Version : 2.

80
CFD Open Series – Annapolis, MD
Adapted & Edited by : Ideen Sadrehaghighi

Turbomachinery of Gas
Turbines in CFD

Unsteady
Unsteady
Flow in
Flow in Axial
Radial
Turbomachines
Turbomachines
(ANSYS)
(ANSYS)
Blade
Interaction
(NUMECA)
Goldman
Annular
Turbine
Cascade
(NASA-TGRID)
1

Table of Contents

1 Introduction ................................................................................................................................ 13
1.1 Preliminaries......................................................................................................................................................... 13
1.2 Full Engine Simulation Methodology .......................................................................................................... 14
1.3 References .............................................................................................................................................................. 15

2 Preliminary Concepts in Rotating Machinery ................................................................. 16


2.1 Vortex....................................................................................................................................................................... 16
2.1.1 Properties of Vortex Flow .................................................................................................................. 16
2.1.1.1 Vorticity ......................................................................................................................................... 16
2.1.1.2 Vortex Types ............................................................................................................................... 16
2.1.1.2.1 Rigid-Body Vortex ............................................................................................... 17
2.1.1.2.2 Irrotational Vortex .............................................................................................. 17
2.1.2 Vortex Geometry .................................................................................................................................... 17
2.1.3 Pressure in Vortex ................................................................................................................................. 18
2.2 Propellers vs Impeller ....................................................................................................................................... 18
2.3 Impeller ................................................................................................................................................................... 19
2.3.1 Types of Impeller ................................................................................................................................... 19
2.3.2 Flow Characteristics for Impeller.................................................................................................... 19
2.3.3 Axial Impellers ........................................................................................................................................ 20
2.3.4 Radial Impellers...................................................................................................................................... 20
2.3.5 Power Number for Impeller .............................................................................................................. 21
2.4 Propeller Aerodynamics................................................................................................................................... 21
2.5 Mixing Tanks ......................................................................................................................................................... 22
2.6 Pumps ...................................................................................................................................................................... 22
2.6.1 Types of Pumps....................................................................................................................................... 23
2.6.2 Axial-Flow Pumps vs. Centrifugal Pumps .................................................................................... 24
2.7 Some Physics on Rotating Disks Flow ........................................................................................................ 24
2.7.1 Experimental Set-Up............................................................................................................................. 24
2.7.1.1 Recirculating Flow .................................................................................................................... 25
2.7.1.2 Instability Flow Patterns ........................................................................................................ 25
2.8 Discussion on Effects Swept and Dihedral Blades ................................................................................. 27
2.9 Linear Cascade ..................................................................................................................................................... 28
2.10 Chapter References ............................................................................................................................................ 29

3 Conservation of Angular Momentum & Rotating Reference Frame ........................ 30


3.1 Flow in Rotating Reference Frame .............................................................................................................. 30
3.1.1 Centrifugal & Coriolis Forces ............................................................................................................ 31
3.1.2 Relative Velocity Formulation .......................................................................................................... 31
3.1.3 Absolute Velocity Formulation......................................................................................................... 32
3.1.4 Early Formulation and Consideration ........................................................................................... 32
3.2 Flows with Rotating Reference Frames ..................................................................................................... 32
3.2.1 Single Rotating Reference Frame (SRF) Modeling ................................................................... 33
3.2.2 Flow in Multiple Rotating Reference Frames (MRF)............................................................... 33
3.2.2.1 Case Study – Mixing Tank ...................................................................................................... 34
3.2.3 The MRF Interface Formulation....................................................................................................... 35
3.2.3.1 Interface Treatment: Relative Velocity Formulation.................................................. 35
3.2.3.2 Interface Treatment: Absolute Velocity Formulation ................................................ 36
2

3.2.3.3 Case Study 1 - Experiments and CFD Calculations on the Performance of a


Non-Reversible Axial Fan ................................................................................................................................ 36
3.2.3.3.1 Introduction........................................................................................................ 36
3.2.3.3.2 Experimental Method ........................................................................................ 37
3.2.3.3.3 Numerical Method ............................................................................................. 38
3.2.3.3.4 Results and Discussion ....................................................................................... 39
3.2.3.3.5 Conclusion .......................................................................................................... 40
3.2.3.4 Case Study 2 - Technologies for the Improvement of Jet fan Installation
Factors 42
3.2.3.4.1 Abstract .............................................................................................................. 42
3.2.3.4.2 Introduction........................................................................................................ 42
3.2.3.4.3 CFD Modelling .................................................................................................... 43
3.2.3.4.4 Technologies Investigated in the Study.............................................................. 43
3.2.3.4.5 CFD Results ......................................................................................................... 44
3.2.3.4.6 Conclusions......................................................................................................... 49
3.2.3.4.7 References .......................................................................................................... 49
3.2.3.5 Case Study 3 - Potsdam Propeller CFD Benchmark .................................................... 50
3.2.3.5.1 Potsdam Propeller CFD Benchmark Description ................................................ 50
3.2.3.5.2 Meshing .............................................................................................................. 50
3.2.3.5.3 TCFD® Simulation Setup ...................................................................................... 52
3.2.3.5.4 TCFD® Simulation Post-processing...................................................................... 53
3.2.3.5.5 Efficiency, Thrust & Torque Coefficient vs. Advance Coefficient ....................... 53
3.2.3.5.6 Conclusion .......................................................................................................... 53
3.2.3.5.7 References .......................................................................................................... 53
3.3 The Mixing Plane Model (MPM) .................................................................................................................... 55
3.3.1 Rotor and Stator Domains .................................................................................................................. 55
3.3.2 The Mixing Plane Concept .................................................................................................................. 56
3.3.3 Mixing Plane Algorithm ....................................................................................................................... 56
3.3.3.1 Mass Conservation Across the Mixing Plane.................................................................. 56
3.4 Sliding Mesh Modeling ...................................................................................................................................... 57
3.4.1 Sliding Mesh Theory ............................................................................................................................. 57
3.4.2 The Sliding Mesh Technique ............................................................................................................. 58
3.4.3 Sliding Mesh Concept ........................................................................................................................... 59
3.4.4 References................................................................................................................................................. 59

4 Elements of Turbomachinery ............................................................................................... 61


4.1 Background............................................................................................................................................................ 61
4.2 Historical Perspectives ..................................................................................................................................... 62
4.3 Modern Turbomachinery as Related to Gas Turbine Engine ............................................................ 62
4.3.1 A Working Cycle Analogy for Gas Turbine Engines ................................................................. 64
4.3.1.1 Working Cycle ............................................................................................................................. 64
4.4 Variance Among Turbojet, Turbofan and Turboprop Engines in Aviation ................................. 66
4.4.1 Turbojet ..................................................................................................................................................... 66
4.4.2 Turbofan .................................................................................................................................................... 67
4.4.2.1 References .................................................................................................................................... 67
4.4.3 Turboprop ................................................................................................................................................. 68
4.4.4 How Does It Work?................................................................................................................................ 69
4.4.5 What is Thrust? ....................................................................................................................................... 69
4.4.6 Afterburner............................................................................................................................................... 70
3

4.4.6.1 References .................................................................................................................................... 71


4.4.7 High Speed Flow Jet Engines ............................................................................................................. 71
4.5 Gas Turbine Performance ................................................................................................................................ 71
4.6 Gas Compressors ................................................................................................................................................. 72
4.6.1 Axial-Flow Compressors ..................................................................................................................... 73
4.6.2 Centrifugal Compressors .................................................................................................................... 73
4.7 Nomenclature of Common Terms ................................................................................................................ 74
4.8 Component of Gas Turbine Engine .............................................................................................................. 77
4.8.1 Inlet .............................................................................................................................................................. 77
4.8.2 Axial Compressor ................................................................................................................................... 77
4.8.3 Diffuser ....................................................................................................................................................... 80
4.8.4 Nozzle ......................................................................................................................................................... 80
4.8.5 Combustor................................................................................................................................................. 81
4.8.6 Axial Gas Turbine ................................................................................................................................... 82
4.9 Variance in Blading Between Compressor and Turbine ..................................................................... 83
4.10 Velocity Triangles in Turbomachines ......................................................................................................... 84
4.11 Energy Exchange with Moving Blades........................................................................................................ 85
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............................................................................................................................................. 92
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
4

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
5.8.9 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 ........................................................................................................ 127
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 Case Study - Interaction of Turbomachinery Components in Large-scale Unsteady
Computations of Jet Engines .................................................................................................................................... 139
6.10.1 Introduction .......................................................................................................................................... 139
6.10.2 Flow solver: SUmb .............................................................................................................................. 140
6.10.3 Results ..................................................................................................................................................... 141
6.10.3.1 Fan / compressor: 20o sector computations ............................................................... 141
6.10.3.2 Turbine: full-wheel computations ................................................................................... 142
6.10.4 Conclusions............................................................................................................................................ 143
6.10.5 References.............................................................................................................................................. 144

7 Rotor-Stator Interaction Treatment (RST) ................................................................... 146


7.1 Physical Perspectives ..................................................................................................................................... 146
7.2 Multi-Passage vs. Multi-Stages ................................................................................................................... 148
7.3 Case for Mixing Plane Model ....................................................................................................................... 148
7.4 Steady Treatment of Interface (Mixing Plane) ..................................................................................... 150
7.4.1 Losses Across the Interface of Mixing Plane ............................................................................ 151
7.4.2 Principles of Flux Conservation .................................................................................................... 152
7.4.3 Case Study 1 - Comparison of Flux Balanced Mixing Models on Q-1.5 Stage Rotor 67
154
7.4.4 Case Study 2 - Modeling of Secondary Flows in Single Blade Rows using Mixing
Plane Approach ....................................................................................................................................................... 156
5

7.4.4.1 Transonic Turbine Stage Meshing and Flow Details ............................................... 156
7.4.5 Case Study 3 - Improvement Methods for Mixing Plane Models ..................................... 158
7.4.5.1 Validation Test Case .............................................................................................................. 159
7.4.5.2 1.5 Stage Transonic Axial Compressor .......................................................................... 159
7.4.6 Frozen Rotor ......................................................................................................................................... 160
7.5 Un-Steady Treatment of Interface............................................................................................................. 161
7.5.1 Sliding Mesh (MRF) ............................................................................................................................ 161
7.5.2 Non-Linear Harmonic Balanced Method (NLHB) .................................................................. 162
7.5.3 Profile Transformation (Pitch Scaling) ...................................................................................... 163
7.5.4 Time Transformation Method (TT) using Phase-Shifted Periodic Boundary
Conditions 164
7.5.5 Revisiting Non-Linear Harmonic Balance (NLHB) Methodology.................................... 165
7.5.5.1 Temporal & Spatial Periodicity Requirement ............................................................ 166
7.5.5.2 Boundary Conditions ............................................................................................................ 166
7.5.5.3 Fourier 'Shape Correction' for Single Passage Time-Marching Solution ........ 168
7.5.5.4 Case Study 1 – 2D Compressor Stage ............................................................................. 169
7.5.5.5 Case Study 2 - 3D Flow in Turbine Cascade ................................................................ 170
7.5.6 Assessment of 2D Steady and Unsteady Adjoint Sensitivities for Rotor-Starter
Interaction 171
7.5.6.1 Case Study ................................................................................................................................. 172
7.5.6.2 Results......................................................................................................................................... 172
7.5.6.2.1 Case 1 - Subsonic Stage .................................................................................... 173
7.5.6.2.2 Case 2 - Transonic Stage ................................................................................... 173
7.5.6.3 Design Sensitivity ................................................................................................................... 173
7.5.6.3.1 Case1 - Subsonic Stage ..................................................................................... 176
7.5.6.3.2 Case2 - Transonic Stage .................................................................................... 176
7.5.6.4 Conclusions ............................................................................................................................... 177
7.6 Case Study - Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design
Perspective...................................................................................................................................................................... 177
7.6.1 Introduction .......................................................................................................................................... 178
7.6.2 Stator-Rotor Interaction in Axial Stages .................................................................................... 179
7.6.2.1 Stator Wake-Rotor Blade Interaction ............................................................................ 180
7.6.2.2 Stator Secondary Lows-Rotor Blade Interaction ...................................................... 181
7.6.2.3 Off-Design Conditions........................................................................................................... 188
7.6.2.4 Stator Shock-Rotor Blade Interaction ............................................................................ 189
7.6.3 Design Perspective ............................................................................................................................. 191
7.6.3.1 Axial Gap .................................................................................................................................... 191
7.6.3.2 End wall Contouring and 3D Blade Geometries ........................................................ 192
7.6.3.3 Cascades Clocking .................................................................................................................. 193

8 Radial Flow ............................................................................................................................... 195


8.1 Centrifugal Compressor................................................................................................................................. 195
8.1.1 Operation Theory................................................................................................................................ 195
8.1.2 Similarities to Axial Compressor .................................................................................................. 195
8.1.3 Components of a simple Centrifugal Compressor ................................................................. 196
8.1.3.1 Inlet .............................................................................................................................................. 196
8.1.3.2 Centrifugal Impeller .............................................................................................................. 196
8.1.3.3 Diffuser ....................................................................................................................................... 197
8.1.3.4 Collector ..................................................................................................................................... 197
8.1.4 Applications........................................................................................................................................... 198
6

8.1.4.1Gas Turbines and Auxiliary Power Units ...................................................................... 198


8.1.4.2Automotive and Diesel Engines Turbochargers and Superchargers ................ 198
8.1.4.3Natural Gas to Move the Gas from the Production site to the Consumer ....... 198
8.1.4.4Oil Refineries, Natural Gas Processing, Petrochemical and Chemical Plants 198
8.1.4.5Air-Conditioning and Refrigeration and HVAC .......................................................... 198
8.1.4.6Industry and Manufacturing to Supply Compressed Air ....................................... 198
8.1.4.7Air Separation Plants to Manufacture Purified End Product Gases .................. 198
8.1.4.8Oil Field Re-Injection of High Pressure Natural Gas to Improve Oil Recovery
199
8.2 Radial Turbine ................................................................................................................................................... 199
8.2.1 Advantages and Challenges ............................................................................................................ 199
8.2.2 Types of Radial Turbines ................................................................................................................. 199
8.2.2.1 Cantilever Radial Turbine ................................................................................................... 200
8.2.2.2 90 Degree IFR Turbine ......................................................................................................... 200
8.2.2.3 Outward-Flow Radial Stages ............................................................................................. 200

9 Best Procedures for Turbomachinery ............................................................................ 202


9.1 Quasi-3D (Q3D) or 3D Simulation ............................................................................................................. 202
9.1.1 2D Simulations ..................................................................................................................................... 202
9.1.2 Quasi-3D (Q3D) Simulation ............................................................................................................ 202
9.1.3 Full 3D Simulations ............................................................................................................................ 202
9.2 Single vs Multi-Stage Analysis .................................................................................................................... 203
9.2.1 Single Stage ............................................................................................................................................ 203
9.2.2 Multi-Stage Analysis .......................................................................................................................... 204
9.2.2.1 Steady Mixing-Plane Simulations .................................................................................... 204
9.2.2.2 Steady Frozen Rotor Simulations .................................................................................... 204
9.2.2.3 Unsteady Sliding Mesh Stator-Rotor Simulations ..................................................... 204
9.2.2.4 Unsteady Harmonic Balance Simulations .................................................................... 204
9.2.2.5 Hybrid Steady-Unsteady Stator-Rotor Simulations ................................................. 205
9.2.2.6 Other Advanced Multi-Stage Methods ........................................................................... 205
9.3 Inviscid or Viscid .............................................................................................................................................. 205
9.4 Transient or Steady-State ............................................................................................................................. 206
9.5 Meshing ................................................................................................................................................................ 207
9.5.1 Mesh Size Guidelines ......................................................................................................................... 208
9.5.2 Case Study - Mesh Resolution Effect on 3D RANS Turbomachinery Flow Simulations
209
9.5.2.1 Formulation of Problems .................................................................................................... 210
9.5.2.2 Conclusions ............................................................................................................................... 210
9.5.3 Boundary Mesh Resolution ............................................................................................................. 211
9.5.4 Periodic Meshing ................................................................................................................................. 211
9.6 Boundary Conditions ...................................................................................................................................... 212
9.7 Turbulence Modeling...................................................................................................................................... 213
9.8 Aero-Mechanics ................................................................................................................................................ 214
9.8.1 Nodal Diameter .................................................................................................................................... 215
9.9 Near Wall Treatment ...................................................................................................................................... 215
9.10 Transition Prediction ..................................................................................................................................... 216
9.11 Numerical Consideration .............................................................................................................................. 216
9.12 Convergence Criteria ...................................................................................................................................... 216
9.13 Single or Double Precision ........................................................................................................................... 217
9.14 Heat Transfer Prediction .............................................................................................................................. 217
7

9.14.1 Keeping it Cool in Gas Turbine ...................................................................................................... 217


9.15 Literature Review and Parallel Processing Tools ............................................................................... 218
9.16 Concluding Remarks ....................................................................................................................................... 219
List of Tables
Table 3.2.1 Mesh Resolution and Number of Cells .............................................................................................. 38
Table 3.2.2 Solution Methods and Boundary Conditions ................................................................................. 40
Table 3.2.3 Number of CFD cells employed in the simulations ...................................................................... 43
Table 3.2.4 The density of air in still conditions can be taken as 1.185 kg m-3. The fan shaft power
was estimated as the product of the blade torque and the rotational speed in radians per second .. 44
Table 3.2.5 Summary of Jetfan Performance ......................................................................................................... 47
Table 3.3.1 Prescribed Boundary zone for Mixing Plane .................................................................................. 56
Table 4.7.1 Glossary of Turbomachinery Terms .................................................................................................. 74
Table 5.8.1 Rotor 37 Optimized Bump Performance Comparison – Courtesy of [John et al.]........ 119
Table 7.1.1 Rotor/Stator Interaction Schemes .................................................................................................. 147
Table 7.4.1 Nomenclature for different Mixing Models Used in Study .................................................... 154
Table 7.5.1 Axial turbine simulation parameters ............................................................................................. 172
Table 9.15.1 Parallel computing tools adopted for parallelization of CFD code used for
turbomachinery – Courtesy of Pinto et al................................................................................................................. 219
List of Figures
Figure 1.2.1 Layout of the KJ66 micro gas turbine .............................................................................................. 14
Figure 1.2.2 Three Types of Full Engine Simulation Methodologies ........................................................... 15
Figure 2.1.1 Vortex created by the passage of an aircraft wing, revealed by colored smoke ............ 16
Figure 2.1.2 Rigid-Body Vortex ................................................................................................................................... 17
Figure 2.1.3 3D Visualization of a Vortex Curve ................................................................................................... 18
Figure 2.1.4 A Plughole Vortex .................................................................................................................................... 18
Figure 2.3.1 Types of Impeller ..................................................................................................................................... 19
Figure 2.3.2 A centrifugal pump uses an impeller with backward-swept arms ..................................... 20
Figure 2.3.3 Flow Direction of Three Different Pumps/Impellers (Courtesy of Global Spec) ........ 20
Figure 2.4.1 Single- (left) and Contra- (right) rotating propeller schematic of the actuator disk
depicting the velocity components ................................................................................................................................ 21
Figure 2.5.1 Axial Flow Impeller (left) and Radial Flow Impeller (right) .................................................. 22
Figure 2.6.1 Centrifugal Pumps ................................................................................................................................... 23
Figure 2.7.1 Sketch of the experimental set-up .................................................................................................... 24
Figure 2.7.2 For s ≥ 0 co-rotation at different speed .......................................................................................... 25
Figure 2.7.3 For s < 0 counter-rotating at different speed .............................................................................. 26
Figure 2.8.1 Geometry of Swept and Dihedral Blades ....................................................................................... 27
Figure 2.8.2 Experimental Pressure Iso-Surfaces ; Left - Without Sweep ;Right - With Forward
Sweep (Courtesy of RÁBAI and VAD [93]) .................................................................................................................. 28
Figure 2.9.1 Linear Turbine Cascade......................................................................................................................... 28
Figure 3.1.1 Rotating Frame of Reference .............................................................................................................. 30
Figure 3.1.2 Centrifugal and Coriolis Force ............................................................................................................ 31
Figure 3.2.1 Single Blade Model with Rotationally Periodic Boundaries................................................... 33
Figure 3.2.2 Mixing Tank Geometry with One Rotating Impeller ................................................................. 34
Figure 3.2.3 Mixing Tank with Two Rotating Impellers .................................................................................... 35
Figure 3.2.4 Interface Treatment for the MRF Model......................................................................................... 35
Figure 3.2.5 An Industrial Axial Fan Used For Ventilation............................................................................... 36
Figure 3.2.6 AMCA 210 test rig [8] ............................................................................................................................. 37
8

Figure 3.2.7 Surface mesh of the rotating zone .................................................................................................... 38


Figure 3.2.8 Cell zones and boundary conditions ................................................................................................ 39
Figure 3.2.9 The streamlines inside the duct at a rotor speed of 1500 rpm............................................. 40
Figure 3.2.10 Torque vs. Volumetric Flow Rate Curves .................................................................................... 41
Figure 3.2.11 CFD Model of Tunnel including Jetfan .......................................................................................... 42
Figure 3.2.12 Mesh for MoJet Bench Thrust Test................................................................................................. 43
Figure 3.2.13 Jetfan with Deflector Vanes ............................................................................................................... 44
Figure 3.2.14 Bench Thrust Test for MoJet (Velocity Contours) ................................................................... 45
Figure 3.2.15 Velocity Contours for Conventional Jetfan (Color Legend as per Figure 3.2.19) ........ 46
Figure 3.2.16 Axial and Circumferential Velocities at Discharge from Conventional Jetfan Silencer
with Struts (33.4° Blade Pitch Angle) ........................................................................................................................... 46
Figure 3.2.17 Velocity Contours with Slanted Silencers (Color Legend as per Figure 3.2.19).......... 48
Figure 3.2.18 Velocity Contours for MoJet with 39° Blade Pitch Angle, with Struts (Color Legend
as per Figure 3.2.19) ............................................................................................................................................................ 48
Figure 3.2.19 Velocity Contours with Deflectors (Color Legend as per 3.2.19) ...................................... 48
Figure 3.2.20 SVA Potsdam Laboratory ................................................................................................................... 50
Figure 3.2.21 Potsdam Propeller Benchmark - POINTWISE Mesh ............................................................... 51
Figure 3.2.22 Domain Decomposition of MRF....................................................................................................... 51
Figure 3.2.23 Potsdam Propeller Benchmark - snappyHexMesh Mesh...................................................... 52
Figure 3.2.24 Surface Static Pressure ....................................................................................................................... 53
Figure 3.2.25 Comparison of TCFD vs. SVA using Pointwise Mesh .............................................................. 54
Figure 3.3.1 Mixing Plane Concepts as Applied to Axial Rotation ................................................................. 55
Figure 3.3.2 Mixing Plane Concepts Applied to Radial Rotation ................................................................... 56
Figure 3.4.1 Illustration of Unsteady Interactions................................................................................................ 57
Figure 3.4.2 Examples of Transient Interaction using Sliding Mesh ............................................................ 58
Figure 3.4.3 Initial position and some translation with Sliding Interface ................................................. 58
Figure 4.1.1 Classification of Turbomachines ....................................................................................................... 61
Figure 4.3.1 Component of Turbomachines and their Thermodynamic (Brayton cycle)
properties ................................................................................................................................................................................. 63
Figure 4.3.2 A comparison between the working cycle of a turbo-jet engine and a piston engine52
....................................................................................................................................................................................................... 64
Figure 4.3.3 The working cycle on a pressure volume ...................................................................................... 65
Figure 4.4.1 Turbojet Engine ........................................................................................................................................ 66
Figure 4.4.2 Low Bypass Turbofan Engine by K. Aainsqatsi ........................................................................... 66
Figure 4.4.3 High Bypass Turbofan Engine by K. Aainsqatsi........................................................................... 67
Figure 4.4.4 Turboprop Engine ................................................................................................................................... 68
Figure 4.4.5 Turboshaft Engine ................................................................................................................................... 68
Figure 4.4.6 A 1D Control Volume around a propulsion system (Courtesy’s of NASA Glen Research
Center) ....................................................................................................................................................................................... 70
Figure 4.4.7 Basic principle of an afterburner ...................................................................................................... 70
Figure 4.4.8 A Turbo/Ram Jet Engine speed of up to Mach 3 and Beyond................................................ 71
Figure 4.6.1 Gas Compressor Types .......................................................................................................................... 72
Figure 4.6.2 Schematics of 3.5-stage of Axial Compressor............................................................................... 73
Figure 4.6.3 A single stage Centrifugal Compressor ........................................................................................... 73
Figure 4.7.1 Blade Related Terminology ................................................................................................................. 77
Figure 4.8.2 Schematic Diagram of fluid properties through an axial compressor stage – Courtesy
of [T. B. Ferguson, Gravdahl, and Egeland] ............................................................................................................... 78
Figure 4.8.3 Pressure and Velocity profile through a Multi-Stage Axial Compressor .......................... 79
Figure 4.8.4 Schematic of a Bell nozzle .................................................................................................................... 81
Figure 4.8.5 Combustor Primary Operating Components ................................................................................ 81
9

Figure 4.8.6 Schematics of Axial Flow Turbine ..................................................................................................... 82


Figure 4.8.7 Turbine Flow Characteristics.............................................................................................................. 82
Figure 4.9.1 Examples of Typical Blades for Compressor and Turbine ...................................................... 83
Figure 4.9.2 Distinguishing Between Compressor and Turbine Blades ..................................................... 84
Figure 4.10.1 Velocity triangles for an Axial Compressor ................................................................................ 84
Figure 4.12.1 Velocity Triangles in Relation to Incident Angle ...................................................................... 87
Figure 4.12.2 Compressor Operating Map .............................................................................................................. 87
Figure 4.13.1 Sample engine Perssure, Velocity and Temperature variation .......................................... 88
Figure 4.13.2 Turbine Inlet Temperature27 ............................................................................................................ 89
Figure 4.14.1 Characteristics Graph of a Compressor ....................................................................................... 90
Figure 4.14.2 Illustration of the Propagation of a Stall Cell in the Relative Frame ................................ 91
Figure 4.14.3 Classical Compressor Surge Cycles ................................................................................................ 91
Figure 5.1.1 Impact of CFD on SNECMA fan performance, over a period of 30 years .......................... 93
Figure 5.3.1 Illustration of S1 and S2 surfaces ...................................................................................................... 95
Figure 5.3.2 Streamline Curvature Method ............................................................................................................ 96
Figure 5.3.3 The turbine Design Process ................................................................................................................. 98
Figure 5.3.4 Flow Path of the Turbine ...................................................................................................................... 99
Figure 5.3.5 Schematics of an airfoil showing stream lines along the radial direction ..................... 100
Figure 5.3.6 3D model of an airfoil showing the passage between adjacent airfoils ......................... 100
Figure 5.4.1 Radial Equilibrium ............................................................................................................................... 100
Figure 5.5.1 Coriolis and Centripetal forces created by the Rotating Frame of Reference.............. 101
Figure 5.6.1 Compression process .......................................................................................................................... 103
Figure 5.6.2 Expansion process ................................................................................................................................ 103
Figure 5.7.1 Linear Turbine Cascade and Computational Domain ............................................................ 104
Figure 5.7.2 A) Default mesh B) Refine mesh............................. 105
Figure 5.7.3 Average Pressure Specification at pressure boundary ......................................................... 106
Figure 5.7.4 Stanton Number Distribution on End-Wall................................................................................ 107
Figure 5.7.5 Stanton number Distribution on Blade Surface for 2D Grid ............................................... 108
Figure 5.8.1 Contours Of Casing Static Pressure Beneath A High-Speed Rotor (550 M/S ............... 109
Figure 5.8.2 Schematic of Shock Structures (A) Datum, (B) S-Shaped Design. Courtesy of [John et
al.].............................................................................................................................................................................................. 111
Figure 5.8.3 Datum Geometry and Optimized Shock Control Bumps on The Mid- Section of Nasa
Rotor 67- From Mazaheri et al.. ................................................................................................................................... 112
Figure 5.8.4 The R37 CFD Domain Used – Courtesy of [John et al.] .......................................................... 113
Figure 5.8.5 Radial Profiles Vs Experimental Data – Courtesy of [John et al.] ...................................... 114
Figure 5.8.6 (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 5.8.7 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 5.8.8 Example 2d CST Bump (Solid Line) And The Four Polynomials Used To Construct It
(Dashed Lines) – Courtesy of [John et al.] ................................................................................................................ 117
Figure 5.8.9 a) Example Individual Bump Geometry ...................................................................................... 117
Figure 5.8.10 Spanwise Slice of the Datum and Optimized R37 Geometries At 60% Span –
Courtesy of [John et al.] .................................................................................................................................................... 118
Figure 5.8.11 Optimized R37 Bump (Blue) Added To The Datum Blade Geometry (Grey) ............ 118
Figure 5.8.12 Datum (Left) And Optimized (Right) Rotor 37 Static Pressure Contours. Flow
Direction Right To Left – Courtesy of [John et al.] ................................................................................................ 119
Figure 5.8.13 Datum (Left) And R37 Optimized (Right) ................................................................................ 120
Figure 5.8.14 Datum (Left) And Optimized (Right) Rotor 37 Separated Flow Contours (Orange).
Flow Direction Right To Left. ......................................................................................................................................... 120
10

Figure 5.8.15 Lift Plots For The Datum and Optimized .................................................................................. 121
Figure 5.8.16 R37 Optimized Characteristic Vs Datum – Courtesy of [John et al.] ............................. 121
Figure 6.1.1 Complex Flow Phenomena Compressors ................................................................................... 123
Figure 6.1.2 Fan Tip Section Geometry ................................................................................................................. 124
Figure 6.2.1 Flow Structures with 5 to 6 Orders of Magnitudes Variations in Length and Time
Scales (LaGraff et al., 2006) ............................................................................................................................................ 125
Figure 6.3.1 Shock Structure in Transonic Fan .................................................................................................. 126
Figure 6.4.1 Pressure Contour of Wake Flow ..................................................................................................... 127
Figure 6.5.1 Instantaneous Absolute Velocity Contour at Nozzle Exit [Matsunuma, 2006] ........... 128
Figure 6.6.1 Unsteady Wakes Convecting in Blade Passage ......................................................................... 129
Figure 6.7.1 Flow Over an Unshrouded Tip Gap ............................................................................................... 130
Figure 6.8.1 Classical Secondary Flow Model ..................................................................................................... 131
Figure 6.8.2 Modern Secondary Flow Model ...................................................................................................... 132
Figure 6.8.3 Vortex pattern of Latest Secondary Flows ................................................................................. 133
Figure 6.8.4 Turbine Secondary Flow Model (Takeishi et al.) .................................................................... 135
Figure 6.9.1 Illustration of formation of hub corner stall together with limiting streamlines and
separation lines ................................................................................................................................................................... 137
Figure 6.9.2 Three-dimensional separations: traditional view and scope of current investigation
.................................................................................................................................................................................................... 137
Figure 6.9.3 Location of monitor point – Courtesy of (Liu et al.) ............................................................... 138
Figure 6.10.1 Schematics of an aircraft jet engine ............................................................................................ 139
Figure 6.10.2 PW 6000 fan/compressor with casing ...................................................................................... 141
Figure 6.10.3 Pressure distribution plotted in log scale ................................................................................ 141
Figure 6.10.4 Instantaneous turbulent kinetic energy distribution in an axial plane showing
wakes and the tip................................................................................................................................................................ 142
Figure 6.10.5 Instantaneous turbulent kinetic energy distribution in a radial plane showing wakes
.................................................................................................................................................................................................... 143
Figure 6.10.6 Entropy distributions for simulations of the first two stages of a modern turbine.
Steady solution with the mixing plane approximation (left), unsteady solution for a 20o sector with
scaled geometry (middle) and unsteady solution for the full wheel (right) .............................................. 144
Figure 7.1.1 Schematics of 3D Concept at IGV/Rotor/Stator Interface ................................................... 146
Figure 7.1.2 Interface Between Rotor/Stator ..................................................................................................... 147
Figure 7.2.1 Difference between Passage and Stages...................................................................................... 148
Figure 7.4.1 Block Computational Domain for a Rotor with guiding Vanes .......................................... 150
Figure 7.4.2 Axial Rotor/Stator Interaction (Schematics Illustrating the Mixing Plane concepts)
.................................................................................................................................................................................................... 150
Figure 7.4.3 A Compressor Pressure Distribution on a Surface using a Mixing Plane ...................... 151
Figure 7.4.4 Schematic of an Artificial Interface Between a Rotor and a Stator (left) and the Virtual
Control Volume Formed by Displacing Two Adjacent Domains (right) ...................................................... 152
Figure 7.4.5 Sketch of Casing Treatment of Rotor 67 (Courtesy of 157) ................................................... 154
Figure 7.4.6 Contour of Relative Mach Number and Iso-Surface of Axial Velocity of Modified Rotor
67............................................................................................................................................................................................... 155
Figure 7.4.7 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.)............................................................................... 155
Figure 7.4.8 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............................................ 157
Figure 7.4.9 Results of the Velocity Contours for a Radial Section at Stator Mid Span using the
Mixing Plane Approach .................................................................................................................................................... 157
Figure 7.4.10 Schematic View of Pitch-Wise Mixing Model .......................................................................... 158
Figure 7.4.11 Instantaneous Distributions at 90% Span. .............................................................................. 159
11

Figure 7.4.12 Total Pressure Calculated by the Frozen Rotor ..................................................................... 160
Figure 7.5.1 Half Stencil and Full Stencil Reconstruction with: A) Intersection, B) Halo-Cell ....... 162
Figure 7.5.2 Relative Velocities Obtained using HB Techniques ................................................................ 163
Figure 7.5.3 Phase shifted Periodic Boundary ................................................................................................... 163
Figure 7.5.4 Phase Shifted Periodic Boundary Conditions............................................................................ 164
Figure 7.5.5 Stagnation Pressure Contours under inlet distortion for NASA Rotor 67 .................... 169
Figure 7.5.6 Instantaneous Pressure Distribution Within the Compressor Stage Using (NLHB) . 169
Figure 7.5.7 Computational Mesh for HB and TRS Methods ....................................................................... 170
Figure 7.5.8 Instantaneous Predictions of Turbulent Viscosity at Mid-Span for HB and TRS
Solutions ................................................................................................................................................................................. 171
Figure 7.5.9 Non-Dimensional Entropy Generation Using Unsteady (HB) vs Steady (MP) ............ 173
Figure 7.5.10 Case2: Non-Dimensional Pressure Contours for the Mixing Plane (MP) simulation
(a) , and Harmonic Balance (HB) at Different Time Instances (bcd) ............................................................ 174
Figure 7.5.11 Mixing Plane vs Harmonic Balance Normalized Entropy Generation Gradients
Obtained with the Adjoint Solution ........................................................................................................................... 176
Figure 7.6.1 Velocity triangles for the free stream (subscript FS) and the wake (subscript W) lows.
V = absolute velocity, W = relative velocity, U = peripheral velocity ............................................................ 180
Figure 7.6.2 Pattern of entropy evolution (bowing, chopping and transport) of the stator wake in
the rotor channel, as foreseen by CFD ....................................................................................................................... 180
Figure 7.6.3 Simplified Schematic of the Secondary Flows System Downstream of a Rotor.......... 182
Figure 7.6.4 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).
.................................................................................................................................................................................................... 183
Figure 7.6.5 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). ........................................................................................................................... 184
Figure 7.6.6 Schematics of the stator vortical structure transport in the rotor passage ................. 184
Figure 7.6.7 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). .................................................................................................... 185
Figure 7.6.8 Relative total pressure coefficient on the whole rotor crown. Experiments at the Fluid
machinery Lab. At Politecnico di Milano (Italy). ................................................................................................... 186
Figure 7.6.9 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).
.................................................................................................................................................................................................... 187
Figure 7.6.10 Standard deviation of the Cptr and δ for the different time frames. Experiments at the
Fluid machinery Lab. at Politecnico di Milano (Italy). ........................................................................................ 188
Figure 7.6.11 Rotor loading effects on the stator-rotor interaction. Experiments at the Fluid
machinery Lab. at Politecnico di Milano (Italy). .................................................................................................... 189
Figure 7.6.12 Vane shock-rotor interaction in axial turbine blades. Red: computation, black:
experiments. Adapted from [Denos et al.]................................................................................................................ 190
Figure 7.6.13 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). .................................................................................................... 191
Figure 7.6.14 Rotor incidence fluctuations in circumferential direction for the different axial gaps
along the blade span.......................................................................................................................................................... 192
Figure 7.6.15 Efficiency trend versus the axial gap. Experiments at the Fluid machinery Lab. at
Politecnico di Milano (Italy). .......................................................................................................................................... 192
Figure 8.1.1 Centrifugal impeller with a highly polished surface likely to improve performance195
12

Figure 8.1.2 Cut-Away View of a Turbocharger showing the Centrifugal Compressor .................... 196
Figure 8.1.3 Jet Engine Cutaway Showing the Centrifugal Compressor among others..................... 197
Figure 8.2.1 Ninety Degree Inward-Flow Radial Turbine Stage ................................................................. 199
Figure 8.2.2 Outward Flow Radial Turbine ......................................................................................................... 200
Figure 9.1.1 Different Flow (2D, Q3D, and full 3D) .......................................................................................... 203
Figure 9.2.2 Full Blade Simulation using Harmonic Balanced Method (Courtesy of CD-adapco) 205
Figure 9.4.1 Transient Blade Row Extensions Enable Efficient Multi-Stage CFD Simulation
(Courtesy of ANSYS.com) ................................................................................................................................................ 207
Figure 9.5.1 Typical Meshing of a Turbomachinery Stage ............................................................................ 208
Figure 9.5.2 Computational Block for a Single Blade ...................................................................................... 208
Figure 9.5.3 Multi-Block Grid for the Space Shuttle Main Engine Fuel Turbine .................................. 209
Figure 9.6.1 Pressure contour plot, 2nd order spatial discretization scheme ...................................... 212
Figure 9.8.1 Analysis provided vibration required for flutter analysis ................................................... 215
Figure 9.8.2 Examples of Nodal Diameter ............................................................................................................ 215
13

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 effects [1].
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,
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
14

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 Wu
[2]. 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
different component analysis tools Figure 1.2.1 Layout of the KJ66 micro gas turbine
(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 1.2.1).
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
15

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 1.2.2). 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 1.2.2 Three Types of Full Engine Simulation Methodologies

1.3 References
[1] Ronald H Aungier, e-Books, “The American Society of Mechanical Engineers”, (ASME.org).
[2] 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.
16

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) [1]. The form plural of vortex is
either vortices or vortexes. Vortices 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 2.1.1). 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 2.1.1 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 mass
[2].
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. 2.1.1
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:
17

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 2.1.2). 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
angular velocity Ω of the fluid around the center of
rotation.

Ω = (0 , 0 , Ω) , r = (x , y , 0) Figure 2.1.2 Rigid-Body Vortex


u = Ω × r = (−Ωy , Ωx , 0) →

ω = ∇ × u = (0, 0, 2Ω) = 2Ω
Eq. 2.1.2
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. 2.1.3
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
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 2.1.3 and can be manufactured as
18

x = t sin(t) , y = t cos(t) , z = t
Eq. 2.1.4
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 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 Figure 2.1.3 3D Visualization of a Vortex Curve
pressure varies as P∞ − K/r , where P∞ is the
2

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 2.1.4). The forward
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 Propellers vs Impeller


Impeller: the rotating part of a centrifugal pump or
mixer, compressor, or other machine designed to move
a fluid by rotation. Impellers for most mixing
applications. Simply put, an impeller is a rotor that
creates an intrinsic sucking force, as part of a pump set
up.
Propeller: a mechanical device for propelling a boat or Figure 2.1.4 A Plughole Vortex
aircraft, consisting of a revolving shaft with two or
more broad, angled blades attached to it. A propeller is used for high shear, high RPM applications. It
is a type of fan that propels fluid by pushing against it, converting rotational force into a linear
motion. (source : Dynamix Agitators & Engineering 360).
19

2.3 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,
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

Figure 2.3.1 Types of Impeller

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.
2.3.1 Types of Impeller
The impeller of a Centrifugal Pump can be of three types as shown in Figure 2.3.1 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 [3].
2.3.2 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 2.3.2). Because
centrifugal pumps are also classified in this manner, the impeller selection depends upon matching
the pump's flow characteristic to that of the impeller [3].
• 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
20

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 engine or by steam driven turbine. The Figure 2.3.2 A centrifugal pump
rotor usually names both the spindle and the impeller when uses an impeller with backward-
swept arms
they are mounted by bolts.
2.3.3 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. (see Figure 2.3.3).

Figure 2.3.3 Flow Direction of Three Different Pumps/Impellers (Courtesy of Global Spec)

2.3.4 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. (see Figure 2.3.3).
21

2.3.5 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 =
n3 D5 ρ
Eq. 2.3.1
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 size [5].

2.4 Propeller Aerodynamics


According to [Gur et al.]1, the left part of Figure 2.4.1 shows a single-rotating propeller with an axial
free stream velocity, V. In this figure the propeller equivalent actuator-disk model is also presented.
The propeller is replaced with an infinitesimal thin actuator-disk which introduces axial and
circumferential momentum to the flow. The additional momentum is resulted with axial and
circumferential induced velocity components, Wa and Wt, respectively. The source of these induced

Figure 2.4.1 Single- (left) and Contra- (right) rotating propeller schematic of the actuator disk
depicting the velocity components

1Ohad Gur , Jonathan Silver, Radovan Dítě and Raam Sundhar, “Contra-Rotating Hover-Propeller Design”, The
61st Israel Annual Conference on Aerospace Sciences, Technion, Haifa.
22

velocities is the propeller thrust and torque. According to Newton’s third law, the thrust and torque
that acts on the propeller is reacted by the same thrust and torque which acts on the flow. Thus, the
air is accelerated to the axial direction due to the thrust and the wake is swirled due to the torque,
i.e. axial and circumferential induced velocities are resulted.
In the current effort, the SR model is extended to a contra-rotating case as appear in right part of
Figure 2.4.1. It shows a contra-rotating propeller and its equivalent actuator-disks model. Each of
the two actuator-disks acts the same manner as a single-rotating case. Main difference is the cross-
induced velocities. Except of the axial and circumferential induced velocities which the disk induces
on itself, the disks induce velocities on each other. The rear disk induces axial cross-induced velocity
on the forward disk, Wa,1on2, and the forward disk induces on the rear disk both axial cross-induced
velocity, Wa,1on2, and circumferential cross-induced velocity, Wt,1on2. Note that the rear disk does not
induce circumferential induced velocity on the forward disk – only the wake swirls. Gur

2.5 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 2.5.1 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
general when there is a deformable interface
to break. Another application of 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-types [4]
Figure 2.5.1 Axial Flow Impeller (left) and Radial
• Propellers Flow Impeller (right)
• Paddles
• Turbines

2.6 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,
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
23

medicine, and as artificial replacements for body parts, in particular the artificial heart and penile
prosthesis. (see Figure 2.6.1),
• 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.

Single Stage Multi Stage

Figure 2.6.1 Centrifugal Pumps

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.
2.6.1 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
❖ Compressed-air-powered double-diaphragm pumps
❖ Rope pumps
• Impulse Pumps
24

➢ Hydraulic ram pumps


• Velocity pumps
➢ Radial-flow pumps
➢ Axial-flow pumps
➢ Mixed-flow pumps
➢ Jet pump
• Gravity Pumps
• Steam Pumps
• Valve less Pumps
2.6.2 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.

2.7 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 disks [6]. 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 fluid2.
2.7.1 Experimental Set-Up
In order to study the flow between
two rotating disks the experimental
set-up shown in Figure 2.7.1 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 2.7.1 Sketch of the experimental set-up
visualization from above and from
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

2 Here the focus is not on the recirculation flow but rather on the instability patterns in rotating fluids.
25

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 glycerol [7]. 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.
2.7.1.1 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.
2.7.1.2 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

Figure 2.7.2 For s ≥ 0 co-rotation at different speed


26

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 2.7.2 (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 2.7.2 (c)). It can
be shown that co-rotation shifts upwards the instability thresholds for circles and positive spirals
[8]. However, threshold line for circles is parallel to the solid body rotation (b = t) indicating that the
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

Figure 2.7.3 For s < 0 counter-rotating at different speed

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 2.7.3 (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 2.7.3 (b). Here negative and positive spirals seem to coexist
without strong interaction, which indicates the difference in their origin. The circles and positive
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 2.7.3(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?
27

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.

2.8 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 Figure 2.8.1 Geometry of Swept and Dihedral Blades
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 2.8.1). 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 flow oppose the classical 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 2.8.2) [9]. Aft sweep has the opposite effects and leads to an
28

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.

Figure 2.8.2 Experimental Pressure Iso-Surfaces ; Left - Without Sweep ;Right - With Forward Sweep
(Courtesy of RÁBAI and VAD [93])

2.9 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
2.9.1). A linear cascade or blade row is built by
setting up prismatic blades with the 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) [10].
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 Figure 2.9.1 Linear Turbine Cascade
constant angular momentum.
29

2.10 Chapter References


[1] Ting , L., & Klein, R. (1991). Viscous Vortical Flows (Lecture Notes in Physics). Springer.
[2] Wikipedia
[3] Presented by: Matt Prosoli, “Centrifugal Pump Overview”, Pumps Plus Inc.
[4] Wikipedia
[5] Engineering 360 powered by IEEE global spec
[6] Miha Meznar, “Fluid Flows In Rotating Frames”, University of Ljubljana, March 2005
[7] 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
[8] Miha Meznar, “Fluid Flows In Rotating Frames”, University of Ljubljana, March 2005.
[9] Gergely 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).
[10] Erik Dick, “Fundamentals of Turbomachines”, Springer Dordrecht Heidelberg New York London.
ISBN 978-94-017-9626-2.
30

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 3.1.1. The origin of the rotating system is located by a position vector r0 (Batchelor,
2000)[1]. 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. 3.1.1
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 [1]. 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 3.1.1 Rotating Frame of Reference


31

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)[2]. 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 3.1.2).

ωω

Figure 3.1.2 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. 3.1.2
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 enthalpy (Hr), also known as the rothalpy [3].
32

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. 3.1.3
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 otherwise [3].
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][4]. 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 & Yocum][5] 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 requirements3. For more complex geometries, it may not be possible to use a single
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)

3 FLUENT 6.3 User's Guide.


33

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 (Figure 3.2.1). 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 (e.g. pressure inlets) there
are restrictions based upon the
velocity formulation which has
been chosen. For additional Figure 3.2.1 Single Blade Model with Rotationally Periodic
information, reader should refer to Boundaries
FLUENT 6.3® user manual.
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)
34

➢ 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
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 model should not be
used, however, if it is 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 3.2.2. (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 3.2.3 shows a geometry that contains two
rotating impellers side by side. This problem would
be modeled using three reference frames: the Figure 3.2.2 Mixing Tank Geometry with One
stationary frame outside both impeller regions and Rotating Impeller
two separate rotating reference frames for the two
impellers. (As noted above, the dashes denote the interfaces between reference frames.)
35

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 Figure 3.2.3 Mixing Tank with Two Rotating Impellers
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 3.2.4). Figure 3.2.4 Interface Treatment for the MRF Model
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 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
36

𝐮 = 𝐮f + 𝛚
⏟×𝐫 + 𝐮t
Swirl Velocity
∇𝐮 = ∇𝐮f + ∇(𝛚 × 𝐫)
Eq. 3.2.1
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.
3.2.3.3 Case Study 1 - Experiments and CFD Calculations on the Performance of a Non-Reversible
Axial Fan
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
CFD simulation results of a six blade axial fan are compared with test data obtained from AMCA
chamber (BACAK, ÜNVERDİ, Salih, 2016)4. 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 3.2.5.
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 3.2.5 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.

4 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

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 before5-6.
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
implementing Latin Hypercube sampling in the design space and multi objective evolutionary
algorithm coupled with response surface approximation surrogate model7-8.
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 3.2.6. 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 3.2.6 AMCA 210 test rig [8]

5 D. Dwivedi, D.S. Dantodiya, “CFD Analysis of Axial Flow Fans with Skewed Blades”, IJETAE, Issue 10, 2013).
6 T. Köktürk, “Design and Performance Analysis of a Reversible Axial Flow Fan”, Master Thesis, METU, 2005.
7 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.
8 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.


38

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 testing 9 the minimum
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, respectively10. 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 3.2.7.
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
the blades. Thin boundary mesh layers required Figure 3.2.7 Surface mesh of the rotating zone
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 Mesh Density Number of Cells
performance and especially the fan noise Coarse Mesh 169477
generation11-12. Convergence of the results are Medium Mesh 178410
tested under grid refinement by generating three Fine Mesh 496616
meshes of different resolution, whose number of
cells are given in Table 3.2.1. Table 3.2.1 Mesh Resolution and Number of
Reynolds averaged continuity and Navier-Stokes Cells
Equations coupled with turbulent kinetic energy

9 AMCA 210: Air Movement and Control Association International, Inc. “Laboratory Methods of Testing Fans for
Aerodynamic Performance Rating “, 1999.
10 M. T. Pascu, “Modern Layout and Design Strategy for Axial Fans”, Ph.D. Thesis at Erlangen University, 2009.
11 Srinivas G., Srinivasa Rao Potti: “Numerical Simulation of Axial Flow Fan Using Gambit and Fluent”. IJRET,

Vol.3, No 3, pp. 586-590, 2014.


12 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.


39

and dissipation rate transport equations of the 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 13. Continuity,
momentum and k-ε turbulence model equations are given below:

∂ρ ∂(ρui )
+ =0
∂t ∂xi
∂ ∂
ρui +
̅̅̅̅ ρu
̅̅̅̅̅
u =
∂t ∂xi i j
∂p ∂ ∂ui ∂uj 2 ∂ui ∂
− + [μ ( + − δij )] + ̅̅̅̅̅
(−ρu ′ u′ )
i j
∂xi ∂xj ∂xj ∂xj 3 ∂xi ∂xj
Eq. 3.2.2
∂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. 3.2.3
∂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. 3.2.4
After generating tetrahedral cells with appropriate resolution above the inflation layers, boundary
conditions are set. In Error! Reference source not found. and Error! Reference source not found. ro
tating and stationary cell zones and boundary conditions are indicated. For more accurate solutions,
second order
discretization is used for
continuity, momentum
and turbulence model
equations.
3.2.3.3.4 Results and
Discussion
Reynolds averaged
continuity, Navier-Stokes
and k-ε turbulence model
equations are solved
simultaneously using
Figure 3.2.8 Cell zones and boundary conditions
Multiple Reference

13 T. Köktürk, “Design and Performance Analysis of a Reversible Axial Flow Fan”, Master Thesis, METU, 2005.
40

Frames, MRF, method. After simulations, validity of the results are checked 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 Turbulence Modeling κ-ε
pressure rise, torque and efficiency vs. Inlet B.C. Velocity Inlet
volumetric flow rate curves are given for Outlet B.C. Pressure Outlet
the three grid resolutions. Difference Blade Pressure Side Periodic
between the calculated and test torque Blade Suction Side Periodic
data is caused by neglecting mechanical Fluid Air
friction in the CFD model.
Fan Speed 1500
Residual For Iteration
30 ∗ Q ∗ ∆p Convergence
1E-5
η=
π∗n∗T Near Wall Treatment Standard Wall Function
Eq. 3.2.5
where Q is the volumetric airflow rate in Table 3.2.2 Solution Methods and Boundary Conditions
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 3.2.10, 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 power and fan rpm.

60 ∗ p
Tmotor =
2∗π∗n
Eq. 3.2.6
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 3.2.9 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 3.2.9 The streamlines inside the duct at a rotor
periodic boundary conditions decreases speed of 1500 rpm
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
41

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 Figure 3.2.10 Torque vs. Volumetric Flow Rate Curves
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.
42

3.2.3.4 Case Study 2 - Technologies for the Improvement of Jet fan Installation Factors
Authors : Fathi Tarada, Karl Else
Affiliations : Mosen Ltd, UK
Originally Appeared : 9th International Conference ‘Tunnel Safety and Ventilation’ 2018, Graz, Austria
3.2.3.4.1 Abstract
Three different technologies for the improvement of Jet fan installation factors, namely deflection
vanes, slanted silencers and MoJets, were calculated using 3D CFD and their performance was
compared to that of a conventional Jet fan. For a fixed installation height of 1.7 m, the MoJet exhibited
an in-tunnel thrust enhancement of 61% above that of a conventional Jet fan, with slightly reduced
power consumption. Slanted silencers increased the installation factor compared to a conventional
Jet fan, but the in-tunnel thrust was reduced due to restrictions on the fan diameter. The deflection
vanes tested in this study were not effective in increasing the in-tunnel thrust, due to the attachment
of the jet to the tunnel floor.
Keywords: tunnel, ventilation, Jet fan, installation factor, efficiency, thrust
3.2.3.4.2 Introduction
Alternative technologies have been proposed for the improvement of Jetfan installation factors,
including deflection vanes (with or without a clearance between the Jetfan outlet and the vanes,
Lotsberg (1997)), the Banana Jet (slanted silencers, Witt and Schütze (2008)) and the MoJet (a
development on the original concept using silencers with an inclined outlet, Tarada (2018)). Each of
these technologies implies different space requirements, power consumption, installation factors
and induced tunnel velocities. Bench thrust tests were simulated for a conventional Jetfan and a
MoJet to establish baseline values for the fan mass flow, power absorbed and axial thrust. The
technologies were then tested in a tunnel environment and the values for the shaft power, axial
thrust, tunnel airflow velocity and installation factor were calculated. The limitations of using the
concept of installation factor in a strongly swirling, 3D flow-field in a tunnel are discussed.

Figure 3.2.11 CFD Model of Tunnel including Jetfan


43

3.2.3.4.3 CFD Modelling


The calculations reported in this paper were undertaken using ANSYS CFX, a commercially available
general-purpose CFD code. A typical CFD model investigated in this study can be seen in Figure
3.2.11. The model represents a 211.6 m long section of tunnel, within which a single Jetfan is located
along the tunnel centerline at the tunnel soffit, 52.8 m from the upstream end. The height of the
tunnel is 6.75 m, and is 16 m wide.
The detailed geometry of the Jetfan hub, blades and motor was included in the same CFD model as
the tunnel geometry. Blade rotation was simulated using the multiple frame of rotation option in
ANSYS CFX, with circumferential averaging. The mass flow through the Jetfans was not prescribed
but was calculated through the CFD code based on a blade rotational speed of 1500 rpm,
corresponding to the speed of a four-pole motor. Reductions in blade rotational speed due to
induction motor slippage were not accounted for. The k-ω SST model of turbulence (Menter, 1993)
was employed, which assisted in identifying any areas of flow separation. The tunnel walls, floor and
soffit were assumed to have a uniform sand roughness height of 0.03 m, giving an equivalent friction
factor (λ = 4f) of 0.024.

Figure 3.2.12 Mesh for MoJet Bench Thrust Test

Boundary layer meshes were attached to all solid surfaces to resolve sharp velocity gradients. Within
the fans and silencers, these boundary layer meshes were calculated via 30-layer prisms with an
initial layer approximately 0.1 mm thick within a Jetfan, and with an expansion factor of 1.2 (Figure
3.2.12). Along the tunnel, the boundary layer mesh was 10-layer with an initial layer 0.3 m thick and
an expansion factor of 1.1. A typical y+ value for the last cell (downstream) along the (rough) tunnel
surfaces was 315, while a typical y+ value on the fan blades and silencers was 15. Table 3.2.3 lists
the number of CFD cells used for each type of technology tested with M = million.
Technology Conventional Deflection Slanted MoJet
Jetfan Vanes Silencers
CFD Cells 25.1 M 72.3 M 25.6 M 27.1 M

Table 3.2.3 Number of CFD cells employed in the simulations

3.2.3.4.4 Technologies Investigated in the Study


The four technologies selected for this study are well established in engineering practice
(conventional Jetfan, deflector vanes and slanted silencers) or represent a recent innovation (MoJet).
The fan, casing, silencer and center body geometries for all four technologies were provided by a
major manufacturer of tunnel ventilation equipment. The same available installation height of 1.7 m
44

was assumed for all four Jetfan technologies. The silencers attached to the fans were selected with
100 mm thickness, and were arranged to be 150 mm below the tunnel soffit.
The Jetfan internal diameter for the conventional Jetfan, deflection vanes and MoJet was set at 1.25
m. A smaller Jetfan diameter (1 m) had to be selected for the slanted silencers (with a 7° angle to the
horizontal) to keep the Jetfan to just within the 1.7 m installation height limit. All four technologies
had “2D” lengths of silencers installed on either side of the fan, where “D” is the internal diameter of
the fan. A mid-range blade pitch angle of 33.4° was selected for the initial calculations with the 1.25
m diameter Jetfans. A blade pitch angle of 39° was also tested for the MoJet, to approximately match

Figure 3.2.13 Jetfan with Deflector Vanes

the shaft power absorbed by the conventional Jetfan. The deflector vanes included a flat section
followed by vanes set at 25° from the horizontal. Following the fan manufacturer’s guidance, the
deflector vanes were installed at a distance of 0.504 m from the silencers at both ends of the Jetfan
(Figure 3.2.13).
3.2.3.4.5 CFD Results
Bench Thrust Results
Bench thrust calculations were Models Conventional Jetfan MoJet
undertaken for the
Blade pitch angle 33.4 33.4
conventional Jetfan and MoJet
options, to provide a baseline Blade torque (Nm) 285.6 275.3
for comparison with Fan shaft power (kW) 44.9 43.2
experimental data. The results Fan mass flow (kg/s) 50.57 51.38
are summarized in Compared Thrust (N) 1759 1815
to measurements, the
calculated conventional Jetfan Table 3.2.4 The density of air in still conditions can be taken as
1.185 kg m-3. The fan shaft power was estimated as the product of
thrust is overstated by around
the blade torque and the rotational speed in radians per second
5% due to the neglect of the
induction motor slippage. The
deviation of the MoJet thrust is expected to be less than 5%, because the reduced motor torque causes
less slippage. (Figure 3.2.14 & Table 3.2.4).
45

Figure 3.2.14 Bench Thrust Test for MoJet (Velocity Contours)

Tunnel Installation Results


The effective thrust of a Jetfan in a tunnel, T, is lower than the bench thrust, due to confining effect of
the tunnel soffit (which reduces the mass flow through the fan) and because of the friction on the
tunnel walls, which is represented by the installation efficiency ηi. The value of T is calculated as:

T = ηi ρAA vA (vA − vT )
Eq. 3.2.7
where AA is the cross section of the Jetfan outlet, vA the jet average velocity and vT the velocity in the
tunnel beyond the direct influence of the Jetfan intake and discharge. The installation factors were
interpreted from the CFD results on the basis of the methodology described by Tarada (2016), as
follows:
∆T
ηi = 1 −
Tmax
Eq. 3.2.8
where
ΔT = increase in skin friction drag above standard case = skin friction drag predicted by CFD –
standard skin friction drag and Skin friction drag in 3D CFD = Sum of predicted wall, soffit and
roadway drag forces Standard skin friction drag = (½) ρv2 TλL(AT/Dh)
where
L = tunnel length (m)
AT = cross-sectional area of tunnel (m2)
Dh = hydraulic tunnel diameter (m)
Since the installation factor relies upon one-dimensional consideration of the tunnel airflow, it may
be of limited value in a tunnel with strongly swirling flow, such as Jetfans with internal struts
removed, as discussed below.
46

Conventional Jetfan
The jet discharged from a conventional Jetfan adheres to the tunnel soffit due to the Coanda effect,
causing a loss of in-tunnel thrust due to friction between the jet and the soffit. When installed with
150 mm clearance to the tunnel soffit, the mass flow through the conventional Jetfan is reduced by

Figure 3.2.15 Velocity Contours for Conventional Jetfan (Color Legend as per Figure 3.2.19)

6.1% from its bench test value, due to the confining effect of the soffit on the inlet silencer. Taking
the reduced mass flow into account, the installation factor calculated from the CFD results is very
close to the value provided by the Kempf correlation (1965):

−1
z 2 z
ηi = [0.0192 ( ) − 0.144 ( ) + 1.27]
DA DA
Eq. 3.2.9
where DA is the outlet diameter of the Jetfan and z denotes the distance between the center axis of
the jet at the outlet and the tunnel wall. This result is somewhat surprising, since Kempf did not

Figure 3.2.16 Axial and Circumferential Velocities at Discharge from Conventional Jetfan Silencer
with Struts (33.4° Blade Pitch Angle)
47

include any swirl in his measurements. The CFD calculations indicated a significant swirl velocity of
up to 10 m/s at the discharge from the silencer following the fan (Figure 3.2.15 & Figure 3.2.16).
Removing the struts holding the fan centre body (which is aerodynamically equivalent to replacing
the struts with drop-rods) increases the outlet swirl, but neither improves the installation factor nor
the tunnel airflow velocity.

MoJet
The MoJet design tested in these calculations comprises silencers with trailing edges that are inclined
at 25° from the vertical. The trailing edge is circular in shape, with a diameter (1.379 m) greater than
that of the fan (1.25 m). convergent/divergent bell mouth is attached to the trailing edge, with a
minimum diameter of 1.332 m. The bell mouth is designed to avoid separation of the flow at the lower
edge of the MoJet inlet silencer, and the same bell mouth is designed to turn the flow at the upper
edge of the MoJet discharge silencer. Since the inlet and outlet areas of the MoJet silencer are
significantly greater than the fan area, there is less resistance to the fan airflow, and the MoJet results
therefore exhibit higher mass flow and less power consumption than conventional Jetfans with the
same blade pitch angle.
The increase in the cross-sectional area of the MoJet silencer leads to a significant pressure recovery
downstream of the fan, and to a reduction in the discharge velocity. Approximately 22% of the kinetic
energy of the flow is recovered as static pressure in the MoJet outlet silencer. This pressure recovery
leads to an increase in the force exerted by the MoJet onto the tunnel air, while the reduction in
discharge velocity reduces the aerodynamic friction between the jet and the soffit. The Coanda effect
is also reduced by the effect of the bell mouth in turning the flow away from the tunnel soffit. These
factors explain the high (near-unity) installation factors reported for the MoJet in Table 3.2.5.

Convention Conventional MoJet MoJet Slanted Jetfan


al Jetfan Jetfan (struts (struts silencers with
removed) removed) deflectors
Blade pitch angle 33.4 33.4 39 39 43 33.4
Tunnel airflow 2.66 2.60 3.62 3.89 2.56 2.26
velocity (m/s)
Installation factor 0.84 0.74 1.00 0.97 0.98 0.60
Fan shaft power 56.7 53.1 56.3 53.7 28.8 56.4
(kW)
Fan mass flow 49.0 50.0 57.4 57.6 33.6 49.4
(kg/s)
% of conventional 100% 91% 161% 156% 86% 74%
Jetfan in-tunnel
thrust

Table 3.2.5 Summary of Jetfan Performance

The MoJet results for a 33.4° blade pitch angle show a 24% increase in the in-tunnel thrust compared
to a conventional Jetfan, with 15% less absorbed power. In order to compare the MoJet with a
conventional Jetfan of equivalent power consumption, the MoJet blade pitch angle was increased
from 33.4° to 39°. This enhances the MoJet in-tunnel thrust to 161% of the conventional Jetfan in-
tunnel thrust, with slightly less power consumption than the conventional Jetfan. Removing the
struts (i.e. replacing them with drop-rods) reduces the installation factor slightly, but installation
factor is an unreliable parameter for prediction of this complex swirling flow-field. Using drop-rods
48

in a MoJet delivers a significant increase in the tunnel air velocity with less power consumption –
arguably the main issues of interest to tunnel designers.

Figure 3.2.17 Velocity Contours with Slanted Silencers (Color Legend as per Figure 3.2.19)

Slanted Silencers
The installation factor is improved by the use of slanted silencers compared to a conventional Jetfan,
but this improvement does not adequately compensate for the loss of thrust due to the reduction of
the Jetfan diameter from 1.25 m to 1 m. This loss in thrust occurred even though the blade pitch angle
was increased to the maximum allowable value without stalling, namely 43°. The in-tunnel thrust
with slanted silencers is reduced to 86% of the conventional Jetfan thrust, with a corresponding

Figure 3.2.18 Velocity Contours for MoJet with 39° Blade Pitch Angle, with Struts (Color Legend as per
Figure 3.2.19)

reduction in tunnel air velocity. The loss of thrust due to a reduction in fan diameter with slanted
silencers would occur in any height-restricted space, and is not a result of the 1.7 m clearance
assumed in this study. However, due to the reduction in fan diameter to 1 m, the thrust/power ratio
with this option is better than any of the 1.25 m diameter fans tested here.

Jetfan with Deflectors


The result with deflectors shows that the swirl in the discharge flow is eliminated, and the flow is
efficiently directed downwards . However, the jet then attaches to the tunnel floor, and significant
friction is generated there. The installation factor is therefore rather poor at only 0.60, and the tunnel

Figure 3.2.19 Velocity Contours with Deflectors (Color Legend as per 3.2.19)
49

air velocity is less than that with a conventional Jetfan. It is recognized that a better result could have
been obtained by deflectors that are better designed. The authors contacted three fan manufacturers
with a request for improved deflector geometry, but no response was received.
The authors originally planned to test the same deflectors coupled directly to the ends of a Jetfan.
However, given the poor performance observed with the detached deflectors, the CFD calculations with
the coupled deflectors were abandoned.
3.2.3.4.6 Conclusions
Each technology that purports to improve the tunnel installation factor can have an effect on the
Jetfan itself (e.g. in terms of reduced or enhanced mass flow), as well as having an effect on the tunnel
air (e.g. by deflecting the jet away from the tunnel soffit, or enhancing the pressure force exerted on
the tunnel air). The most promising technology identified by this study has been the MoJet, which can
increase the in-tunnel thrust by 61% above the conventional Jetfan value, with a slightly reduced
power consumption. A research project is underway at the Institute of Aerodynamics at RWTH
Aachen University to manufacture 1:18 scale models of Jetfans with different silencer geometries
including slanted silencers and MoJets using 3D printing, and to measure their tunnel installation
factors. Full-scale tests of MoJets in a tunnel are also planned.
3.2.3.4.7 References
Kempf, J. (1965), Einfluss der Wandeffekte auf die Treibstrahlwirkung eines Strahlgebläses,
Schweizerische Bauzeitung, 83. Jahrgang, Heft 4, Seiten 47-52.
Lotsberg, G. (1997), Investigation of the Wall-friction, Pressure Distribution and the Effectiveness of Big
Jetfans with Deflection Blades in the Fodnes Tunnel in Norway, 9th International Symposium on
Aerodynamics and Ventilation of Vehicle Tunnels, Aosta Valley, Italy.
Menter, F. R. (1993), Zonal Two Equation k-ω Turbulence Models for Aerodynamic Flows, AIAA Paper
93-2906.
Tarada, F. (2016), Innovation in Jetfan Design, 7th International Symposium on Tunnel Safety and
Security, Montreal.
Tarada, F. (2018), Optimized Tunnel Ventilation Device, International PCT Patent Application Number
PCT/GB2018/000029.
Witt, K.C. and Schütze, J. (2008), Effective Thrust Transformation inside Tunnels with Jet Fans (Banana
Jet), 4th International Conference ‘Tunnel Safety and Ventilation’, Graz.
50

3.2.3.5 Case Study 3 - 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.
3.2.3.5.1 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
velocity field measurements at different operation Figure 3.2.20 SVA Potsdam
Laboratory
conditions. A detailed description of the open water tests
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 - Figure 3.2.20). 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. 3.2.10
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.
3.2.3.5.2 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].
51

Figure 3.2.21 Potsdam Propeller Benchmark - POINTWISE Mesh

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. (Figure 3.2.21).
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 (Figure 3.2.22). 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 3.2.22 Domain Decomposition of MRF


52

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
front, or violating quality criteria. If an element stops advancing this did not prevent adjacent

Figure 3.2.23 Potsdam Propeller Benchmark - snappyHexMesh Mesh

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 M14 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.
3.2.3.5.3 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.

14 M= million
53

3.2.3.5.4 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. (Figure 3.2.24).

Figure 3.2.24 Surface Static Pressure

3.2.3.5.5 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 (Figure 3.2.25).
3.2.3.5.6 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.
3.2.3.5.7 References
[1] Barkmann, U., Potsdam Propeller Test Case (PPTC) - Open Water Tests with the Model Propeller
VP1304, Report 3752, Schiffbau Versuchsanstalt 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, Schiffbau Versuchsanstalt 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.
54

[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

Figure 3.2.25 Comparison of TCFD vs. SVA using Pointwise Mesh


55

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 Error! Reference source not found. and
Figure 3.3.2. 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 3.3.2 and Error! Reference source not found.) 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

Figure 3.3.1 Mixing Plane Concepts as Applied to Axial Rotation


56

conformal; that is, the nodes on the stator exit 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
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 3.3.2 and
Error! Reference source not found., p
rofiles of averaged total pressure (p0), Figure 3.3.2 Mixing Plane Concepts Applied to Radial
direction cosines of the local flow angles in Rotation
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, tangential, and axial directions (αr ; αt ; αz), are
Up Stream Down Stream
computed at the stator inlet and used as a boundary
condition on the rotor exit. Passing profiles in the manner Pressure Outlet Pressure Inlet
described above assumes specific boundary condition Pressure Outlet Velocity Inlet
types have been defined at the mixing plane interface. The Pressure Outlet Mass flow Inlet
coupling of an upstream outlet boundary zone with a
downstream inlet boundary zone is called a mixing plane Table 3.3.1 Prescribed Boundary
pair. In order to create mixing plane pairs, the boundary zone for Mixing Plane
zones must be as prescribed as
Table 3.3.1.
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
57

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
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 3.4.1, and
generally classified as follows:
• Potential interactions:
flow unsteadiness due to
pressure waves which
propagate both upstream
and downstream.
• 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 Figure 3.4.1 Illustration of Unsteady Interactions
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 3.4.2 (a) and Figure 3.4.2 (b), you must use sliding meshes. If you are
interested in a steady approximation of the interaction, you may use the multiple reference frame
58

model or the mixing plane model, as described before.

(a) Rotor-Stator
Interaction (b) Blower

Figure 3.4.2 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

(b) Sliding Mesh


(a) Initial Position

(c) Rotor /Starter


Interactions (d) Linear Grid Interface

Figure 3.4.3 Initial position and some translation with Sliding Interface

Figure 3.4.3
59

be positioned so that it has fluid cells on both sides. For example, the grid interface for the geometry
shown in Figure 3.4.3-(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 3.4.3 (a-d)).
Figure 3.4.3-(a) and Figure 3.4.3-(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 3.4.3-(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 3.4.3 (a). (You will need to change these wall zones to some other appropriate
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.
In the example shown in Figure 3.4.3 (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.
3.4.4 References
[1] Batchelor, G. (2000). An Introduction to Fluid Dynamics (Cambridge Mathematical Library). Cambridge:
Cambridge University Press. doi:10.1017/CBO9780511800955
[2] David Apsley, “Fluid-Flow Equations”, spring 2017.
[3] FLUENT 6.3 User's Guide
[4] Johannes Vassiliou Soulis, “An Euler Solver For Three-Dimensional Turbomachinery Flows”, International
Journal For Numerical Methods In Fluids, Vol. 20,L-30, 1995.
[5] 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.
60
61

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)15. A common feature of
these devices is that they all work with fluid and have rotating component. Gorla16 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 4.1.1 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 4.1.1 Classification of Turbomachines

15 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.
16 R. S. R. Gorla. Turbomachinery: Design and Theory. CRC Press, 2003.
62

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 17.

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 . 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 in18. 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

17 See previous.
18 M. T. Schobeiri. Turbomachinery: Flow Physics and Dynamic Performance. Springer, Berlin, 2005. 2, 37, 38
63

Figure 4.3.1 Component of Turbomachines and their Thermodynamic (Brayton cycle) properties

unsteady flow are becoming common place. Automation of CFD, coupling with 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 4.4.2
illustrates of a low bypass 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)19.

19 From Wikipedia, the free encyclopedia.


64

4.3.1 A Working Cycle Analogy for Gas Turbine Engines


Another descriptive for gas turbine is offered by (Rolls-Royce, 1986)20 which might be an interesting
for some users.
1 - Gas turbine engine is essentially a heat engine using air as a working fluid to provide thrust. To
achieve this, the air passing through the engine has to be accelerated; this means that the velocity or
kinetic energy of the air is increased. To obtain this increase, the pressure energy is first of all
increased, followed by the addition of heat energy, before final conversion back to kinetic Energy in
the form of a high velocity jet efflux.
4.3.1.1 Working Cycle
2 - The working cycle of the gas turbine engine is similar to that of the four-stroke piston engine.
However, in the gas turbine engine, combustion occurs at a constant pressure, whereas in the piston
engine it occurs at a constant volume. Both engine cycles (Figure 4.3.2) show that in each instance
there is induction, compression, combustion and exhaust. These processes are intermittent in the

Figure 4.3.2 A comparison between the working cycle of a turbo-jet engine and a piston engine52

case of the piston engine whilst they occur continuously in the gas turbine. In the piston engine only
one stroke is utilized in the production of power, the others being involved in the charging,
compressing and exhausting of the working fluid. In contrast, the turbine engine eliminates the three
’idle’ strokes, thus enabling more fuel to be burnt in a shorter time; hence it produces a greater power
output for a given size of engine.

20 The Jet Engine, 5th ed. , Rolls Royce, ISBN 0 902121 2 35


65

3 - Due to the continuous action of the turbine engine and the fact that the combustion chamber is
not an enclosed space, the pressure of the air does not rise, like that of the piston engine, during
combustion but its volume does increase. This process is known as heating at constant pressure.
Under these conditions there are no peak or fluctuating pressures to be withstood, as is the case with
the piston engine with its peak pressures in excess of 1,000 lb/in2. It is these peak pressures which
make it necessary for the piston engine to employ cylinders of heavy construction and to use high
octane fuels, in contrast to the low octane fuels and the light fabricated combustion chambers used
on the turbine engine.
4 - The working cycle upon which the gas
turbine engine functions is, in its simplest
form, represented by the cycle shown on
the pressure volume diagram in Error! R
eference source not found.. Point A
represents air at atmospheric pressure
that is compressed along the line AB.
From B to C heat is added to the air by
introducing and burning fuel at constant
pressure, thereby considerably
increasing the volume of air. Pressure
losses in the combustion chambers (Part
4) are indicated by the drop between B
and C. From C to D the gases resulting
from combustion expand through the Figure 4.3.3 The working cycle on a pressure volume
turbine and jet pipe back to atmosphere. Diagram52
During this part of the cycle, some of the
energy in the expanding gases is turned into mechanical power by the turbine; the remainder, on its
discharge to atmosphere, provides a propulsive jet.
5 - Because the turbo-jet engine is a heat engine, the higher the temperature of combustion the
greater is the expansion of the gases. The combustion temperature, however, must not exceed a value
that gives a turbine gas entry temperature suitable for the design and materials of the turbine
assembly.
6 - The use of air-cooled blades in the turbine assembly permits a higher gas temperature and a
consequently higher thermal efficiency.
66

4.4 Variance 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
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.

Figure 4.4.1 Turbojet Engine

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 4.4.1).

Figure 4.4.2 Low Bypass Turbofan Engine by K. Aainsqatsi


67

4.4.2 Turbofan
The turbofan or fanjet is a type of airbreathing jet engine that is widely used in aircraft propulsion21.
The word "turbofan" is a portmanteau of "turbine" and "fan": the turbo portion refers to a gas turbine
engine which achieves mechanical energy from combustion [1] and the fan, a ducted fan that uses the
mechanical energy from the gas turbine to force air rearwards. Thus, whereas all the air taken in by
a turbojet passes through the combustion chamber and turbines, in a turbofan some of that air
bypasses these components. A
turbofan thus can be thought of as a
turbojet being used to drive a ducted
fan, with both of these contributing to
the thrust.
The ratio of the mass-flow of air
bypassing the engine core to the
mass-flow of air passing through the
core is referred to as the bypass
ratio (BPR). The engine produces
thrust through a combination of these
two portions working together;
engines that use more jet thrust
relative to fan thrust are known as
low-bypass turbofans (see Figure
4.4.2), conversely those that have
considerably more fan thrust than jet Figure 4.4.3 High Bypass Turbofan Engine by K. Aainsqatsi
thrust are known as high-bypass
(Figure 4.4.3). Most commercial aviation jet engines in use today are of the high-bypass type and
most modern military fighter engines are low-bypass [2][3]. Afterburners are not used on high-
bypass turbofan engines but may be used on either low-bypass turbofan or turbojet engines. Modern
turbofans have either a large single-stage fan or a smaller fan with several stages. An early
configuration combined a low-pressure turbine and fan in a single rear-mounted unit.
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 (Figure 4.4.3).
4.4.2.1 References
[1] Marshall Brain (April 2000). "How Gas Turbine Engines Work". howstuffworks.com.
Retrieved 2010-11-24.
[2] Hall, Nancy (May 5, 2015). "Turbofan Engine". Glenn Research Center. NASA. Retrieved October
[3] Michael Hacker; David Burghardt; Linnea Fletcher; Anthony Gordon; William Peruzzi (March 18,
2009). Engineering and Technology. Cengage Learning. p. 319. ISBN 978-1-285-95643-5.
Retrieved October 25, 2015. All modern jet-powered commercial aircraft use high bypass turbofan
engines

21 Wikipedia
68

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 4.4.4). 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 4.4.4 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 4.4.5). 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 4.4.5 Turboshaft Engine
engine generate more power. The
69

turbine of a turbocharger is driven by engine exhaust gasses, and a supercharger is similar but is
directly powered by the engine22.
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
speed; the power output is enough to drive the compressor, combustion is continuous, and the starter
system can be disengaged.
A short video from YouTube explains better this concepts. https://youtu.be/KjiUUJdPGX0
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 frame23.
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

22 Aviation weekly.
23 University of Cambridge, Compressor and Turbine stages.
70

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 4.4.6) as

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. 4.4.1

Figure 4.4.6 A 1D Control Volume around a propulsion system (Courtesy’s of NASA Glen Research
Center)
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
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 velocity24.
4.4.6 Afterburner
Conferring to Wikipedia, an afterburner (or reheat in British English) is an additional combustion
component used on some jet engines, mostly those on military supersonic aircraft. Its purpose is to
increase thrust, usually for supersonic flight, takeoff, and combat. Afterburning injects
additional fuel into a combustor in the jet pipe behind (i.e., "after") the turbine, "reheating" the
exhaust gas (Figure 4.4.7). Afterburning significantly increases thrust as an alternative to using a
bigger engine with its attendant weight penalty,
but at the cost of very high fuel consumption
(decreased fuel efficiency) which limits its use to
short periods. This aircraft application of reheat
contrasts with the meaning and implementation
of reheat applicable to gas turbines driving
electrical generators and which reduces fuel
consumption [1]. Jet engines are referred to as
operating wet when afterburning is being used Figure 4.4.7 Basic principle of an afterburner

24 NASA – Glen Research Center.


71

and dry when not [2]. An engine producing maximum thrust wet is at maximum power, while an
engine producing maximum thrust dry is at military power [3].
4.4.6.1 References
[1] Gas Turbine Design, Components and System Design Integration, Meinhard T. Schobeiri, ISBN 978
3 319 58376 1, p. 12/24
[2] Ronald D. Flack (2005). Fundamentals of jet propulsion with applications. Cambridge, UK:
Cambridge University Press. ISBN 0-521-81983-0.
[3] Graham, Richard H. (July 15, 2008). Flying the SR-71 Blackbird: In the Cockpit on a Secret
Operational Mission. MBI Publishing Company. p. 56. ISBN 9781610600705.
4.4.7 High Speed Flow Jet Engines
The turbo/ram jet engine (Figure 4.4.8) combines the turbo-jet engine (which is used for speeds up
to Mach 3) with the ram jet engine, which has good performance at high Mach numbers (Rolls-Royce,
1986). The engine is surrounded by a duct that has a variable intake at the front and an afterburning
jet pipe with a variable nozzle at the rear. During takeoff and acceleration, the engine functions as a

Figure 4.4.8 A Turbo/Ram Jet Engine speed of up to Mach 3 and Beyond

conventional turbo-jet with the afterburner lit; at other flight conditions up to Mach 3, the
afterburner is inoperative. As the aircraft accelerates through Mach 3, the turbo-jet is shut down and
the intake air is diverted from the compressor, by guide vanes, and ducted straight into the
afterburning jet pipe, which becomes a ram jet combustion chamber. This engine is suitable for an
aircraft requiring high speed and sustained high Mach number cruise conditions where the engine
operates in the ram jet mode.

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
72

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
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

Axial
Dynamic
Centrifugal Single Acting

Reciprocating Double Acting


Compressor Types
Diaphram

Positive Displacement Vane

Scroll

Rotery Liquid Ring

Screw

Lobe

Figure 4.6.1 Gas Compressor Types


73

main types of gas compressors are illustrated in Figure 4.6.1. 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
4.6.2). 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 Figure 4.6.2 Schematics of 3.5-stage of Axial
Mach number. Beyond about 5 stages or a 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.
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 Figure 4.6.3 A single stage Centrifugal
are also used in internal combustion engines as Compressor
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 4.6.3).
74

4.7 Nomenclature of Common Terms


Before going further, it is prudent to get familiarize our self with terminology used in industry
regarding components of turbomachines25. From personal experience, it is an important issue. Some
of these are shown in Figure 4.7.1 and shown alphabetically in
Table 4.7.1 below.

Table 4.7.1 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.
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.

25David 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

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 ratio of the hub radius to the tip radius
radio
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
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
76

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 instruments
(conditions) 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
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
77

Figure 4.7.1 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 engine26. 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
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,

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


78

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 the blade flow paths. The
stator vanes slow the air by
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
process is repeated fourteen
times as the air flows from
the first stage through the
fourteenth stage. Figure
4.8.2 shows one stage of the
compressor and a graph of
the pressure characteristics
through the stage. (see
[Niazi]27).
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 4.8.2.
Alternatively, assuming
incompressible, constant Figure 4.8.2 Schematic Diagram of fluid properties through an axial
compressor stage – Courtesy of [T. B. Ferguson, Gravdahl, and
density, and with no body
Egeland]
force, we can use Bernoulli’s

27Saeid Niazi, “Numerical Simulation of Rotating Stall and Surge Alleviation in Axial Compressors”, A Thesis
Presented to the Academic Faculty, Georgia Institute of Technology, 2000.
79

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 velocity28. 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 4.8.3. 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
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

Figure 4.8.3 Pressure and Velocity profile through a Multi-Stage Axial Compressor

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

28 MIT OpenCourseWare
80

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.29 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.)30. 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.
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 rapid31. 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 Error! Reference s
ource not found.). Such a nozzle flow occurs at entry to all turbomachines and in the stationary blade

29 MIT, OpenCourseWare.
30 S. L.
Dixon, “Fluid Mechanics and Thermodynamics of Turbomachinery”, 5th edition, Senior Fellow at University
of Liverpool, 1978-1998.
31 See 13.
81

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 Figure 4.8.4 Schematic of a Bell nozzle
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
4.8.5). The secondary air controls the flame pattern, cools the liner walls, dilutes the temperature

Figure 4.8.5 Combustor Primary Operating Components

of the core gasses, and provides mass. This cooling air is critical, as the flame temperature is above
82

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 4.8.7 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
energy is converted to mechanical energy. Velocity, temperature, and pressure of the gas are

Figure 4.8.6 Schematics of Axial Flow Turbine


83

sacrificed in order to rotate the turbine to generate shaft power. Figure 4.8.7 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 4.8.6. 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
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. A short video from YouTube explains better this concepts.
https://youtu.be/KjiUUJdPGX0

4.9 Variance in Blading Between Compressor and Turbine


There is quite a difference between Compressor and Turbine blading. Aside from shape of it, they
are number of stages and arrangement of it. While Compressor blades are generally thin and straight,
and resemble a tiny rectangular wing with low camber thickness. Turbine 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 4.9.1. To distinguish between high pressure and low pressure stages (compressor or

Typical Compressor Blade ( Air


Defence Museum) Typical Turbine Blade

Figure 4.9.1 Examples of Typical Blades for Compressor and Turbine

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. (Figure
4.9.2).
84

Compressor
• Area increase: pressure rise
• Flow deceleration: thick boundary layers
• Little flow turning: many stages

Turbine
• Area decrease: pressure drop
• Flow acceleration: thin boundary layers
• Large flow turning: few stages
Figure 4.9.2 Distinguishing Between Compressor and Turbine Blades

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 (Figure
4.10.1). 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 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 Figure 4.10.1 Velocity triangles for an Axial Compressor
by β. The component of the velocity
85

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 4.10.1).

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
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 Turbomachinery32
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. 4.11.1
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. 4.11.2
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. 4.11.3
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. 4.11.4
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. 4.11.4 it is obvious that:

32 “Compressor and turbines”, Aero students.


86

• 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. 4.11.5
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
equation are what we call absolute frame velocities (as opposed to relative frame velocities).33 It is
given fact that:
• If angular momentum increases across a blade row, then T 2 > 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)

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 have34. 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. 4.12.1
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 4.12.1 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
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 4.12.1, this requires

33MIT, OpenCourseWare.
34 Breuer, B., Bissinger, N., C., “Encyclopedia of Aerospace Engineering – Volume 8 - Chapter EAE 573-Basic
Principles – Gas Turbine Compatibility – Gas Turbine Aspects”.
87

Figure 4.12.1 Velocity Triangles in Relation to Incident Angle

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 4.12.1 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
4.12.2, 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. Figure 4.12.2 Compressor Operating Map
88

Further increasing the pressure ratio is equivalent to further rise of airfoil aerodynamic loading and
losses increase due to 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 provided35. 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 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 operate36. 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

Pressure Temperature Velocity

Figure 4.13.1 Sample engine Perssure, Velocity and Temperature variation

35 See 71.
36 “Fundamentals of Gas Turbine Engines”, Cast-Safety.org.
89

adding higher cost for the direct cooling of the first turbine stage airfoils and other components.
Figure 4.13.1 shows the temperature, velocity and pressure variation across a gas turbine engine37.
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 4.13.2). 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

Figure 4.13.2 Turbine Inlet Temperature27

slightly decreased. It is also known that the power consumption of the cool inlet air is of considerable
concern since it decreases the net power output of the gas turbine38-39. 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

37 Shahrokh Sorkhkhah, “Gas Turbine Fundamentals”, Faculty of Karaj , Azad University Design Director of Iran
Gas Turbine Company.
38 Amjed Ahmed Jasim AL-Luhaibi, Mohammad Tariq, “Thermal Analysis of Cooling Effect on Gas Turbine Blade”,

eISSN: 2319-1163 | pISSN: 2321-7308.


39 Posted by: Abdul Nassar, “Optimizing the Cooling Holes in Gas Turbine Blades”, SoftInWay® Incorporated,

2016.
90

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.

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 on40. 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. 4.14.1
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
4.14.1. 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 Figure 4.14.1 Characteristics Graph of a Compressor
initial and final temperatures are
T04 and T05.
4.14.1 Stall
Let’s examine the air entering the rotor41. Previously, we have assumed that this air has exactly the

40 MIT OpenCourseWare.
41 See previous.
91

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 4.14.2).
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 spreads
around the rotor. And it does Figure 4.14.2 Illustration of the Propagation of a Stall Cell in the
this opposite to the direction of Relative Frame
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 4.14.1).
4.14.2 Compressor Surge42
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 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 Figure 4.14.3 Classical Compressor Surge Cycles
the whole compressor to start vibrating at a high

42 MIT OpenCourseWare.
92

frequency (see Figure 4.14.3). 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]43.
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 flow44. It occurs, when the flow reaches supersonic velocities. Choked flow can
also occur at the compressor. If we look at the right side of Figure 4.14.1, 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 (Figure 4.1.1). 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]45, or Introduction to wind turbine Aerodynamics (Springer 2014)
among others. For Steam and Hydraulic turbines, an excellent references are given by Wikipedia.

43 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.
44 See previous.
45 Magdi 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 transfer46, 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
5.1.1 shows the impact of CFD
on SNECMA fan performance
over a 30 year period. The flows
in blade passages may be
laminar, turbulent and Figure 5.1.1 Impact of CFD on SNECMA fan performance, over a
transitional, and may include period of 30 years
wake flow, and 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”47 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

46 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
47 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 field48. 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 5.1.1 exhibits the impact of CFD on the performance improvement
of aircraft engine in SNECMA (France) over a period of almost 30 years49. 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 points50.

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 Wu51 , 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

48 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.
49 J. F. Escuret, D. Nicoud, and Ph. Veysseyre,”Recent advances in compressors aerodynamic design and analysis”,

AVT TP/1, RTO/NATO, 1998.


50 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.
51 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
5.3.1). 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 Figure 5.3.1 Illustration of S1 and S2 surfaces
by a single 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 in52. The stream surface represented by

s(r, ψ, z) = 0
Eq. 5.3.1
As depicted in Figure 5.3.2. 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 5.3.1). 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 predicted53.

Figure 5.3.2 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

52 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.
53 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 distribution54. 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 mentioning55.
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 decisions56. 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 5.3.3. 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

54 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.
55 Ian K. Jennions, “Elements of a Modern Turbomachinery Design System”, GE Aircraft Engines, One Neumann

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


56 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 5.3.3 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 surface57. 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

57 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 5.3.4. 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 5.3.4.

Figure 5.3.4 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 5.3.5 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 5.3.6. To ensure slope and curvature smoothness of the geometry,

Figure 5.3.6 3D model of an


Figure 5.3.5 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 5.4.1 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. 5.4.1
Writing dm = ρ r dϴ dr and ignoring terms of 2nd order we obtain:

1 dp cθ2
=
ρ dr r
Eq. 5.4.2
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 5.4.1 Radial Equilibrium
ρ dr ρ dr dr dr
1 dp 0 dc
= cx x + θ
c d
(rcθ )
ρ dr dr r dr
Eq. 5.4.3
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. 5.5.1
Introducing this into the mass conversation and after some manipulation we obtain,

∂r ρ
+ ∇. (ρ𝐰) = 0
∂t
Eq. 5.5.2
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. 5.5.3
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 5.5.1 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 5.5.1 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)
M omentum: + + 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. 5.5.4
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]58. 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:

V d t dV +  F − G dA = VS dV where W = ρ , ρu , ρE and


dW
v = u − rω
T

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


Eq. 5.5.5
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 in59.

5.6 Efficiency Effects in Turbomachinery60


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

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

Eq. 5.6.1
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 5.6.2 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
For Expansion : η tt = = = 01 02
ideal change in energy Δh os h 01 − h 02s
Eq. 5.6.2
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 5.6.1 Compression process Figure 5.6.2 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. 5.6.3
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 61-62.

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]63. 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 5.7.1). 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 5.7.1 Linear Turbine Cascade and Computational Domain

61 S.L. Dixon, B.Eng., PH.D., “Fluid Mechanics, Thermodynamics of Turbomachinery”, Senior Fellow at the
University of Liverpool, UK.
62 Damian Vogt, “Efficiencies”, Turbomachinery Lecture Notes, 2007.
63 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 5.7.2 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 5.7.2 A) Default mesh B) Refine mesh


106

shown in Figure 5.7.2 B. The block boundaries of the structured 3-block mesh and the boundary
conditions used are highlighted in the Figure 5.7.1. 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 5.7.2(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 5.7.2(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 5.7.2). 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 value64. 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 5.7.3 Average Pressure Specification at pressure boundary
averaged pressure, and f = 1
recovers the specified pressure. The results of this improvement displayed in Figure 5.7.3.
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

64 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 5.7.4). 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 5.7.4 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 5.7.5 for the v2-f and SA model,
respectively.

Figure 5.7.5 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 5.7.5. At this stage it is not clear whether this is due to the specification of the inlet
109

conditions or the turbulence model65.

5.8 Case Study 2 - Using Shock Control Bumps To Improve Transonic Compressor
Blade Performance66
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.]67. 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 5.8.1 Contours Of Casing Static Pressure Beneath A High-Speed Rotor (550 M/S
Tip Speed) With Pronounced Negative Camber. From Prince – Courtesy of [Prince]

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

Fan/Compressor Blade Performance”, GT2018-77065.


67 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 efficiency68.
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 shock69.
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]70 designed a rotor with pronounced negative camber (see Figure 5.8.1). 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]71 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.]72 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 5.8.2. 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.

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

of Aircraft, 17(1), pp. 28–37.


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

109(3), pp. 340–345.


72 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 5.8.2 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.]73 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]74 carried out numerical studies into shock control in
1987, describing the behavior of shock-induced separation. [Sommerer et al.]75 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.]76 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.]77 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

73 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.
74 Drela, M., and Giles, M. B., 1987. “Viscous-inviscid analysis of transonic and low Reynolds number airfoils”. AIAA

journal, 25(10), pp. 1347–1355.


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

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


76 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.


77 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]78, 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 5.8.3.

Figure 5.8.3 Datum Geometry and Optimized Shock Control Bumps on The Mid- Section of Nasa Rotor
67- From Mazaheri et al..

The work by [Mazaheri & Khatibirad]79 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

78 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.
79 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 5.8.4. 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 5.8.4 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]80. Recent work has also been carried out by [Chima]81 and [Hah]82. 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 5.8.5 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 5.8.5 Radial Profiles Vs Experimental Data – Courtesy of [John et al.]

5.8.6 Flow Field For the Datum Case


Figure 5.8.6 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 5.8.6-b and the orange contour of zero axial velocity in Figure 5.8.6-a. At the point where
the shock impinges on the suction surface, its interaction with the boundary layer causes it to

80 Denton, J., 1997. “Lessons from rotor 37”. Journal of Thermal Science, 6(1), pp. 1–13.
81 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.
82 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 83. 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 5.8.6 (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 5.8.7 shows the flow features of the
blade design as the flow rate is varied (see [John et al.]84. 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

83 Alistair John, Ning Qin, and Shahrokh Shahpar, “Using Shock Control Bumps To Improve Transonic
Fan/Compressor Blade Performance”, GT2018-77065.
84 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 5.8.7 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 Adjoint85 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

85 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.]86
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 5.8.8 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 5.8.8 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 5.8.9 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

86Alistair 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 5.8.9.
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 5.8.10 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 5.8.10. A slice at 60% span is shown. The 3D Figure 5.8.11 Optimized R37 Bump
geometry compared to the datum is shown in Figure (Blue) Added To The Datum Blade
Geometry (Grey)
5.8.11.
119

The bump applied to the datum geometry varies radially, 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 5.8.12. 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 5.8.14. 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 5.8.1 Rotor 37 Optimized Bump Performance Comparison – Courtesy of [John et
here) and al.]
the datum
in Table 5.8.1. It can be seen that the efficiency benefit is greatest for the continuous bump design.

Figure 5.8.12 Datum (Left) And Optimized (Right) Rotor 37 Static Pressure Contours. Flow Direction
Right To Left – Courtesy of [John et al.]

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.]87. 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

87John, 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 5.8.13 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 5.8.13 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 5.8.15 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 5.8.14 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 5.8.16. 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 5.8.15 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.]88.

Figure 5.8.16 R37 Optimized Characteristic Vs Datum – Courtesy of


[John et al.]

88John, 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 6.1.1, the flow fields of fan designs are complex and highly three-
dimensional, and almost always unsteady. The flow-path hub contour shown in Figure 6.1.1
suggests significant radial velocity components, especially at the fan entrance and strong interactions

Figure 6.1.1 Complex Flow Phenomena Compressors

between the shock, 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 6.1.2
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
124

• Sets the maximum flow rate.


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 6.1.2 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,)89. 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

89Lakshminarayana, 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

flows, for instance film cooling flows nearby the cooling holes. The complexity is mainly reflected in
the following areas:

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 6.2.1 with broad spectrum. According to Greitzer90, 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 6.2.1 Flow Structures with 5 to 6 Orders of Magnitudes Variations in Length and Time
Scales (LaGraff et al., 2006)

90 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 6.3.1 Shock Structure in Transonic Fan


127

6.3.1 Interactions in Transonic Fan


The shock structure associated with transonic fans is complicated by the 3D nature of the flow field
and operating range over which the fan must operate91. Figure 6.3.1 (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 6.4.1 (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 6.6.1 by [Smith, 1966; Stieger
& Hodson, 2005]. This unsteady transport process Figure 6.4.1 Pressure Contour of Wake
could last to the next few blade rows and mix with new Flow
wake flows 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.

91 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.
128

[Matsunuma]92 investigated this interaction effect on a low speed turbine of single stage, with the
instantaneous absolute velocity contour pattern at the nozzle exit shown in Figure 6.5.1. 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 6.5.1 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

92T.Matsunuma, “Unsteady flow field of an axial-flow turbine rotor at a low Reynolds number”, ASME-GT06,
number 90013, Spain, 2006.
129

(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 6.6.1):
• 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 6.6.1 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 6.7.1 (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
130

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
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

Figure 6.7.1 Flow Over an Unshrouded Tip Gap

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
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 as93:
• 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

93 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 flow94. 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 6.8.1 (a-b), is proposed by

(a) Classical View


(Hawthorne, 1955)

(b) Secoundary Losses in


presence of secoundary
vortex flow in classical
view

Figure 6.8.1 Classical Secondary Flow Model

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

Hawthorne95 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 6.8.1 (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]96 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 6.8.2 (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.]97 who Figure 6.8.2 Modern Secondary Flow Model
proposed the well-known

95 W. R. Hawthorne,” Rotational flow through cascades part 1: the components of vorticity.” Journal of Mechanics
and Applied Mathematics, 8(3):266–279, 1955.
96 A. Fritsche. Str¨omungsvorg¨ange in schaufelgittern. Technische Rundschau Sulzer, 37(3), 1955.
97 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 6.8.3 (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
twofold98: by Langston et (a) Sharma and Butler, 1987
al99 who proposed the
well-known modern
vortex model in cascade.
Three vortices are
presented in this model,
as depicted in Figure
6.8.3 (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
twofold100:
Figure 6.8.3 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
Sieverding101 in 1977. While, the results of this experiment also show the counter-rotating vortex,

98 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.
99 L. S. Langston, “Three-dimensional flow within a turbine blade passage”, Journal of Engineering for Power,

99(1):21–28, 1977.
100 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.
101 P. Marchal and C. H. Sieerding, “Secondary flows within turbomachinery blading’s”, CP 214, AGARD, 1977.
134

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]102 proposed a secondary flow pattern which is slightly different to that
from Langston. This pattern, shown in Figure 6.8.3 (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]103 and [Sieverding]104. However, in 1988, another
pattern is given by [Goldstein and Spores]105, shown in Figure 6.8.3 (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]106 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 Wang107. 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 Jilek108 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]109. 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]110 reviewed these new models after the [Sieverding’s] review. Laster in the same year,
[Zhou and Han]111 gave a more comprehensive review of all these models. They concluded that the
good 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

102 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.
103 J. Moore and A. Ransmayr, “Flow in a turbine cascade part 1: losses and leading edge effects”, ASME, 1983.
104 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.


105 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.


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

ASME-GT86, number 170, 1986.


107 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.


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

ASME-GT86, number 170, 1986.


109 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.


110 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.


111 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

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 compressors112. 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 6.8.4 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 6.8.4 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 6.8.4, 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
side flows is shown as the attachment line in Figure 6.8.4. It intersects with the separation
line at the saddle point.

112N W Harvey, “Some Effects of Non-Axisymmetric End Wall Profiling on Axial Flow Compressor Aerodynamics.
Part I: Linear Cascade Investigation”, Proceedings of GT2008.
136

• 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 6.8.4,
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]113:
• 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 6.9.1. where the illustration of formation of hub corner stall together
with limiting streamlines and separation lines, (Lei et al.).
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

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


137

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 expected to be at
least 20% of 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
Figure 6.9.1 Illustration of formation of hub corner stall together
small variations in leading edge
with limiting streamlines and separation lines
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]114. The detrimental effects of
these separations have historically
been predicted by correlations based
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 Figure 6.9.2 Three-dimensional separations: traditional
leading edge, which are characteristic view and scope of current investigation

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

of an eroded blade. 3D
separations always occur on
compressor blades in the
corner between the suction
surface and the end wall 115
(Figure 6.9.2). 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.
Another study of 3D corner
separation has been done by
(Liu et al.)116. It was claimed
that the turbulence model,
Delayed Detached Eddy
Simulation (DDES) method has Figure 6.9.3 Location of monitor point – Courtesy of (Liu et al.)
better ability than RANS to
capture the corner separation. Figure 6.9.3 displays the unsteady character of corner separation
region with help of 4 monitoring positions (P1 - P4) in the mainstream.

115 Gbadebo, S. A., Cumpsty, N. A., and Hynes, T. P., 2005, “Three-Dimensional Separations in Axial
Compressors,” ASME J. Turbomachinery , 127, pp. 331–339.
116 Yangwei Liu, Hao Yan, Lipeng Lu, “Investigation of Corner Separation in a Linear Compressor Cascade Using

DDES”, Proceedings of ASME Turbo Expo 2015: Turbine Technical Conference and Exposition GT2015, June 15
– 19, 2015, Montréal, Canada, GT2015-42902.
139

6.10 Case Study - Interaction of Turbomachinery Components in Large-scale


Unsteady Computations of Jet Engines
Authors : Georgi Kalitzin, Gorazd Medic,Edwin van der Weide, and Juan J. Alonso
Affiliation : Stanford University, Stanford, CA 94305-4035
Citation : Georgi Kalitzin, Gorazd Medic, Edwin van der Weide and Juan Alonso. "Interaction of
Turbomachinery Components in Large Scale Unsteady Computations of Jet Engines”, AIAA 2007-
519. 45th AIAA Aerospace Sciences Meeting and Exhibit. January 2007.
6.10.1 Introduction
The objective of the Stanford ASC project [2] is to develop a framework able to perform multi-
disciplinary, integrated simulations on massively parallel platforms [3,4]. This paper focuses on the
turbomachinery computation and, in particular, on the physics of interaction of different
turbomachinery components in the engine. Typical flow features such as tip and horse-shoe vortices
as well as blade wakes will be discussed for these multi-component turbomachinery simulations. The
compressor and turbine of a modern turbofan engine, Figure 6.10.1, typically have two counter-
rotating concentric shafts to allow for different rotational speeds of their components as well as a

Figure 6.10.1 Schematics of an aircraft jet engine

reduction of net torque. The low-pressure parts rotate at a lower rate than the high-pressure
components. Typical rotation rates are 5,000 to 7,000 RPM for the former and 15,000 to 20,000 RPM
for the latter. The compressor and turbine themselves consist of a series of rotors and stators for
which the blade counts are normally chosen such that no sector periodicity occurs. Combined with
the inherently unsteady nature of turbomachinery flows due to the motion of the rotors, the full
wheel geometry needs to be considered in a time accurate numerical simulation of the flow.
The computational requirements for such a simulation are severe. The high-pressure compressor
(HPC) alone consists of 5 stages (rotor/stator combinations) and 50 to 200 blade passages per stage.
Since approximately a million nodes are required per blade passage to obtain a grid-converged
Reynolds-Averaged Navier-Stokes (RANS) solution, the computational mesh for a full wheel HPC
140

simulation contains 500 million to 1 billion nodes. The turbine consists of less stages due to the
favorable pressure gradient. However, a full wheel simulation still requires 150 million to 300 million
nodes. The spatial mesh is to be integrated in time for 2,000 to 10,000 time steps, based on the
estimate that 50 to 100 time steps are needed to resolve a blade passing, to remove the transient
effects. Alone the full wheel unsteady HPC simulation will require 20 to 40 million CPU hours on
today’s fastest computers. Adding the low-pressure compressor (LPC), fan as well as high and low-
pressure turbine (HPT and LPT, respectively), the computational requirements are far beyond what
is currently affordable for practical applications and therefore approximations are used to reduce the
computational costs.
The most widely-used industrial practice for solving turbomachinery problems is the mixing plane
assumption[1]. A circumferential averaging of the flow variables is applied at the interface between
rotor and stator. These average quantities are then imposed as upstream and downstream values for
the following and preceding blade rows respectively and a steady-state computation can be
performed for both the rotor and the stator. Due to this averaging and the periodicity assumption
only one blade passage needs to be simulated per blade row, independently of the blade counts.
Although this assumption models the mean effect of the rotor/stator interaction, all the unsteady
information is lost due to the averaging.
An alternative approach used to perform an unsteady simulation is to choose a periodic sector of, for
example 20o, where the blade counts are changed such that the full wheel can be split into 18 sections
and periodicity conditions can be used. The pitch and chord of the blades then need to be adjusted to
preserve the flow blockage. Because of these changes it is clear that only approximate information
can be obtain from an unsteady sector simulations. Nevertheless, as we intend to show, interesting
results about the interactions between different turbomachinery components can still be obtained.
6.10.2 Flow solver: SUmb
Despite the progress made in unstructured grid technology during the last 10 years, the quality of
solutions obtained on structured grids is still superior compared to their unstructured counterparts.
This is especially true for high Reynolds number RANS simulations where high aspect ratio cells must
be used to capture the anisotropic flow phenomena. In combination with the fact that it is relatively
straightforward to create multi-block structured grids for the geometries used in the turbomachinery
components of a jet engine, all the compressor and turbine simulations carried out within the
Stanford ASC program use multi-block structured grids.
The flow solver performing these simulations is SUmb, which has been developed under the
sponsorship of the Department of Energy Advanced Strategic Computing (ASC) Initiative. SUmb
solves the compressible Euler, laminar Navier-Stokes and RANS equations on multi-block structured
meshes; SUmb is a parallel code, suited for running on massively parallel platforms [5].
SUmb can be used to solve steady-state problems, unsteady problems (with moving geometries) and
time periodic problems, for which a Fourier representation is used for the time derivative leading to
a coupled space-time problem [6,7]. A number of turbulence models is available, including k-ω [8]
Spalart-Allmaras [9] and v2-f [10]. To reduce the grid resolution requirements in the near wall region
adaptive wall functions are used [11, 12].
For the discretization of the inviscid fluxes either a central difference scheme augmented with
artificial dissipation [13] or an upwind scheme in combination with Roe’s approximate Riemann
solver [14] is used. The viscous fluxes are computed using a central discretization. All the results
presented in this paper are obtained with a second order cell-centered discretization. The second
order implicit time integration scheme is used for all unsteady computations. The resulting nonlinear
system is solved using the dual time-stepping approach [15]. The convergence is accelerated via a
standard geometrical multi-grid algorithm in combination with an explicit multi-stage Runge-Kutta
scheme.
141

6.10.3 Results
This section presents some preliminary results for unsteady flow computations of both the
fan/compressor and the turbine components of a typical aircraft jet-engine. Both the compressor
and the turbine grids have the same topology consisting of an O-grid around the blade and an H-type
grid in the passage. The tip gap regions of the rotors have also been resolved. The initial spacing for
both the turbine and the fan/compressor grids corresponds to an average y+ value of 60. The total
number of cells in a passage is
approximately 350,000 for the turbine.
For the compressor grid the number of
cells per passage differs per blade row,
but on average approximately 500,000
cells per passage are used. The subsonic
inflow and outflow boundary conditions
correspond to the regular cruise
condition of the engine. For all results
shown the turbulence is modeled using
the k-ω model using wall functions.
6.10.3.1 Fan / compressor: 20o sector
computations
For the 20o sector simulations the blade
counts are changed so that the full wheel
can be split into 18 sections and
periodicity conditions can be used. The
pitch and chord of the blades are adjusted
to preserve the flow blockage. The steady Figure 6.10.2 PW 6000 fan/compressor with casing
solutions obtained with the mixing plane
assumption are used as initial conditions
for the unsteady computations. The
computational grid consists of 204 blocks
and 57 million cells (the second level grid
has about 7 million cells); 1,200
processors were used on the LLNL ALC
machine for the fine grid computations
and 400 processors were used for the
second level grid. The physical time step
is chosen so that a blade passage of the
blade row with the highest blade count is
resolved with 50 time steps. For the low-
pressure components (fan, LPC and LPT)
this corresponds to 2,700 time steps per
revolution, while for the high-pressure
components (HPC and HPT), which rotate
at a higher speed, 6,300 time steps per
revolution must be taken.
The computational geometry is shown in
Figure 6.10.2. The airflow through the Figure 6.10.3 Pressure distribution plotted in log scale
fan is split into a part that goes through
the LPC and a part that goes through the bypass. The latter is not included in these computations; an
exit bypass pressure boundary condition is specified.
142

The fan / compressor simulation is far more challenging to compute than the flow through the
turbine due to the large adverse pressure gradient. The pressure at the high-pressure compressor
exit is about 30 times larger the pressure at the fan inlet, see Figure 6.10.3 (note that a log scale is
used for the pressure plotted here).
The flow is pushed through the compressor by the work transmitted from the rotating blades to the
fluid and an accurate modeling of the boundary layers is crucial to achieve the designed pressure
ratio. In the computations, the flow tends to reverse its direction as soon as large pockets of
separation occur in the passage.
As a consequence, a special algorithm had to be developed to create an initial flow field. For this, the
flow is first computed on a very coarse grid - here we use the 3rd multigrid grid, i.e. a grid that is 64
times coarser than the finest grid. The computation is carried out by first slowly raising the rotational
speed of the wheels while keeping the pressure low at the exit, and then by increasing the pressure
once the full rotational speed was achieved. Although the 3rd multigrid level does not resolve the
boundary layers at all, it is quite remarkable that even on
this grid a pressure increase of about 65% of the design
pressure ratio could be achieved. The solution is then
interpolated on the next finer grid and the procedure of
slowly raising the back pressure is continued. Performing
computations on successively refined grids also gives
insight into the grid convergence and the order of
accuracy of the approach.
Results after 10000 time steps computed on the second
level grid are presented in Figure 6.10.4 and Figure
6.10.5; the turbulent kinetic energy is plotted in an axial
cross-section just downstream of the fan and on a surface
at a certain radial distance from the hub (corresponding
to mid-span in the LPC), respectively. Clearly visible are
the wakes of the blades in the fan/LPC, as well as the tip
vortex from the fan blades. The large wakes from the fan
blades are preserved over many stages of the low
pressure compressor. These wakes may amplify
turbulence in the LPC creating acoustic noise. The Figure 6.10.4 Instantaneous turbulent
interaction of these wakes with the downstream stators kinetic energy distribution in an axial
and rotors and its effects on the efficiency and flow plane showing wakes and the tip
capacity in the low-pressure compressor are currently vortex
investigated in detail. Specifically, the comparisons are
being made to the current industry practices to compute the flow through the low pressure
compressor, which usually don’t model these unsteady wakes originating from the fan blades.
6.10.3.2 Turbine: full-wheel computations
Due to the smaller number of blade rows in the turbine, the computational cost for the unsteady
simulation of the full wheel turbine are smaller than for the compressor. Consequently most of the
attention for the full wheel simulations has been paid to the high-pressure turbine augmented with
one stage of the low pressure part. The computational grid consists of 496 blocks and 88 million cells.
The disk space needed to store this grid in double precision is 2.1 GBytes. In addition, at least 6 Gbytes
of disk space is needed to store the set of independent variables in the solution file. However, if more
information is stored in the solution file this number increases. As for the 20o sector simulation, one
revolution of the HPT is resolved with 2,700 time steps.
143

This case has been run


on the LLNL ALC
machine on either 300,
600, 1,200 or 1,800
processors, depending
on the available
resources; SUmb writes
the solution in a single
file and a restart can be
made on a different
number of processors
due to the fully
integrated parallel
preprocessor. The
solution has been
advanced 600 time
steps starting from the
mixing plane solution. A
comparison with the
mixing plane and
unsteady 20 sector
o
Figure 6.10.5 Instantaneous turbulent kinetic energy distribution in a
simulation is shown in radial plane showing wakes
Figure 6.10.6. The
quantity displayed is
entropy on a surface located half-way between the hub and the casing. For the unsteady simulations
an instantaneous distribution is shown. The difference between the steady computation using the
mixing plane assumption and the unsteady simulations is evident. The unsteady nature of the flow
field combined with the different blade counts leads to a strong interaction (both downstream and
upstream) between the rotor and stator. This information is lost due to the circumferential averaging
and results in a completely different picture of the flow field. The difference between both unsteady
simulations is not as clear, but subtle differences can be distinguished. Due to the rescaling of the
blade counts in the 20o sector simulation the wake patterns of the preceding blade rows differ from
the true geometry and hence the interaction between the blade rows shows different frequencies.
6.10.4 Conclusions
An unsteady simulation of the flow in the turbomachinery provides a significant improvement
compared with the widely-used industrial practice of steady flow computation using the mixing plane
assumption. The unsteady nature of the flow field leads to a strong interaction between the rotor and
stator wakes, and on a larger scale, it leads to a strong interaction between the components of the
turbomachinery. In the mixing plane assumption, this information is lost due to the circumferential
averaging resulting in a completely different picture of the flow field.
The unsteady flow simulation carried out for a 20o sector of a scaled fan/low-pressure
compressor/high pressure compressor ensemble has shown that the wakes of the fan blades are
preserved over many stages of the low pressure compressor. The interaction of these wakes with the
downstream turbomachinery stages effects the performance of the low-pressure compressor and
may well contribute to noise generation. Although, as shown for the turbine, the unsteady flow
144

Figure 6.10.6 Entropy distributions for simulations of the first two stages of a modern turbine. Steady
solution with the mixing plane approximation (left), unsteady solution for a 20o sector with scaled
geometry (middle) and unsteady solution for the full wheel (right)

results for a full wheel and a 20o sector simulation differ significant less when compared to the steady
flow computation using the mixing plane assumption, the effect of rescaling of the blade counts
remains to be investigated. The different blade counts effects the wake pattern and interaction
frequency. Nevertheless, unsteady computations of a scaled sector provide a relatively cheap way
towards understanding the physics of unsteady turbomachinery flows.
Acknowledgments
We thank the U.S. Department of Energy for the support under the Advanced Simulation and Computing
Program. We also thank Pratt & Whitney for providing the engine geometry, computational meshes,
helpful comments and discussions.
6.10.5 References
[1] Denton, J. and Singh, U., “Time Marching Methods for Turbomachinery Flows,” VKI LS 1979-07,
1979.
[2] Center for Integrated Turbulence Simulations, “Annual ASC Report,” http://cits.stanford.edu,
2006.
[3] Schl¨uter, J., Wu, X., van der Weide, E., Hahn, S., and Alonso, J., “Integrated LES- RANS of an Entire
High- Spool of a Gas Turbine,” AIAA paper 06-0897, 44th AIAA Aerospace Sciences Meetings and
Exhibit, Reno, NV, January 2006.
[4] Medic, G., You, D., Kalitzin, G., van der Weide, E., Alonso, J., and Pitsch, H., “Integrated RANS-LES
Computations of an Entire Gas Turbine Jet Engine,” AIAA, 45th AIAA Fluid Dynamics Conference and
Exhibit, Reno, NV, January 2007.
[5] E. van der Weide, G. Kalitzin, J. S. and Alonso, J., “Unsteady Turbomachinery Computations Using
Massively Parallel Platforms,” AIAA paper 06-0421, 44th AIAA Aerospace Sciences Meetings and
Exhibit, Reno, NV, January 2006.
145

[6] Gopinath, A. and Jameson, A., “Time Spectral Method for Periodic Unsteady Computations over
Two- and Three-Dimensional Bodies,” AIAA paper 05-1220, AIAA 43rd Aerospace Sciences Meeting
and Exhibit, Reno, NV, January 2005.
[7] van der Weide, E., Gopinath, A., and Jameson, A., “Turbomachinery Applications with the Time
Spectral Method,” AIAA paper 05-4905, 35th AIAA Fluid Dynamics Conference and Exhibit, Toronto,
Ontario, June 2005.
[8] Wilcox, D., “Reassesment of the scale-determining equation for advanced turbulence models.”
AIAA Journal , Vol. 26, 1988, pp. 1299–1310.
[9] Spalart, P. and Allmaras, S., “A one-equation turbulence model for aerodynamic flows.” La
Recherche Aerospatiale, Vol. 1, pp. 1–23.
[10] Durbin, P., “Separated flow computations with the k-"-v2 model.” AIAA Journal , Vol. 33, 1995.
[11] Kalitzin, G., Medic, G., Iaccarino, G., and Durbin, P., “Near-wall behavior of RANS turbulence
models and implications for wall function,” Journal of Computational Physics, Vol. 204, 2005, pp.
265–291.
[12] Medic, G., Kalitzin, G., and van der Weide, E., “Adaptive wall functions with applications.” AIAA
paper 06-3744, 36th AIAA Fluid Dynamics Conference and Exhibit, 2006.
[13] Jameson, A., Schmidt, W., and Turkel, E., “Numerical solution of the Euler equations by finite
volume methods using Runge Kutta time stepping schemes,” AIAA paper 81-1259, 1981.
[14] Roe, P., “Fluctuations and Signals - A Framework for Numerical Evolution Problems,” Numerical
Methods for Fluid Dynamics, Academic Press, 1982.
[15] Jameson, A., “Time Dependent Calculations Using Multigrid, with Applications to Unsteady
Flows Past Airfoils and Wings,” AIAA paper 91-1596, June 1998.
146

7 Rotor-Stator Interaction Treatment (RST)


7.1 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 flows117. 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

Figure 7.1.1 Schematics of 3D Concept at IGV/Rotor/Stator Interface

117Xiaodong 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.
147

of the time integration method118. Figure 7.1.1 shows


Steady Un-Steady
a schematics of 3D point of view in either case. All of
Mixing Plane Sliding Mesh
these approximations can be considered to be
different variations of reduced-order models. In Frozen Rotor Harmonic Balance
order to use the same solver, the flows in stator and Time Transformation
rotor should be calculated in the stationary frame of
Table 7.1.1 Rotor/Stator Interaction
reference and the rotating frame of reference,
Schemes
respectively. However, a critical problem is how to
transfer the information downstream and upstream
at the interface of stator and rotor. The quality of the flow predictions for multistage turbomachinery
strongly depends on the treatment of rotor/stator interaction. Figure 7.1.2 illustrates the interface
between them. Two general approaches as steady and unsteady interactions are available as detailed

Figure 7.1.2 Interface Between Rotor/Stator

in Table 7.1.1.

Arathi K. Gopinath, Edwin van der Weidey, Juan J. Alonsoz, Antony Jamesonx, Stanford University, Stanford,
118

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”
148

7.2 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
7.2.1. 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 7.2.1 Difference between Passage and Stages
shape, therefore, in principle, a
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
row119.

7.3 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.]120 , [Pengcheng
& Fangfei]121 , and [Wang]122. 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]123 , 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]124,
employs an interface, called mixing plane, between adjacent blade rows, where the circumferentially

119 Turbomachinery Simulation, ANSYS blog.


120 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.
121 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.
121 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.


121 Denton J. D. “The calculation of three dimensional viscous flow through multistage turbomachines”, 1990.
121 See 148.
122 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.
123 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.
124 Denton J. D. “The calculation of three dimensional viscous flow through multistage turbomachines”,1990.
149

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]125 , 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 :
➢ 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.

125Adamczyk 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.
150

7.4 Steady Treatment of Interface (Mixing Plane)


The simplest treatment of R/S interface is the
stage or Mixing Plane method proposed by
[Denton]126. 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 7.4.2. 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 problem127. At some prescribed
iteration interval, the flow data at the mixing Figure 7.4.2 Block Computational Domain for a
plane interface are averaged in the Rotor with guiding Vanes
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 of either the axial or
the radial coordinate, 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 7.4.1 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, tangential, and
axial directions (αr, αt, αz), are
computed at the stator inlet
and used as a boundary Figure 7.4.1 Axial Rotor/Stator Interaction (Schematics Illustrating
condition on the rotor exit. the Mixing Plane concepts)
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 plane128 can be used to overcome

126 J. D. Denton, “The calculation of three-dimensional viscous flow through multistage Turbomachinery”,
Journal of Turbomachinery, 114(1):18–26, 1992.
127 Release 12.0 © ANSYS, Inc. 2009-01-22.
128 NUMECA International, Brussels, “Fine/Turbo User Manual V8”, October 2007.
151

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. Unfortunately it doesn’t
capture the whole physics. This is usually
evident by visual inspection of in interface
(mixing) plane as an imaginary line
between the cascades. Figure 7.4.3
displays a compressor Pressure
Distribution on a surface at constant radius Figure 7.4.3 A Compressor Pressure Distribution on a
half way between the hub and the casing Surface using a Mixing Plane
using a Mixing Plane computation.
7.4.1 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 rows129. 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
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]130 and
[Pullan]131 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 in132,

129 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.
130 Fritsch, G., and Giles, M. B., “An Asymptotic Analysis of Mixing Loss,” ASME J. Turbomachines, 1995.
131 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.


132 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.


152

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.
7.4.2 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 7.4.4133. 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
segment should be equal to those into the domain on the other side of the segment. This can be easily

Figure 7.4.4 Schematic of an Artificial Interface Between a Rotor and a Stator (left) and the Virtual
Control Volume Formed by Displacing Two Adjacent Domains (right)
understood through forming a virtual control volume, the hatched part as shown in Figure 7.4.4, 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 7.4.4, 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

Ding Xi Wang, “An Improved Mixing-Plane Method for Analyzing Steady Flow Through Multiple-Blade-Row
133

Turbomachines”, Siemens Industrial Turbomachinery Ltd.,


Waterside South, Lincoln LN5 7FD, UK.
153

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. 7.4.1
The area of the segmental surface is given by

S = ∫ dsrdθ
0
Eq. 7.4.2
Dividing Eq. 7.4.1 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. 7.4.1/Eq. 7.4.2, should be conserved. It should be noted the
circumferential area averaged fluxes as defined in Eq. 7.4.1/Eq. 7.4.2 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
differences, 3 methods is been proposed by [Wang]134. 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 135.

134 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.
135 See Previous.
154

7.4.3 Case Study 1 - Comparison of Flux Balanced Mixing Models on Q-1.5 Stage Rotor 67
To study the performance of flux balanced mixing models on simulating the complex flow, such as
shock wave and reversed flow on the interface, the quasi-1.5-stage Rotor 67 with small geometric
modification is then
investigated . An artificial step
of about 6% span is imposed
on the casing contour ahead of
the second interface to
produce reversed flow on the
interface. The designed and
modified casing contours are
shown in Figure 7.4.5. The
grid with the same topology
and cell number used for flow
solver validation is generated.
Table 7.4.1 lists the
nomenclature for Mixing
models for steady
computation. For a listing of
relative deviations of flow
variables across the 1st and Figure 7.4.5 Sketch of Casing Treatment of Rotor 67 (Courtesy of
157)
2nd interfaces of the modified
quasi-1.5 stage Rotor 67 vs. the
circumferential velocity, see [YaLu et al.]136. 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 Steady Mixing Model Nomenclature
the flux balanced models. By the flux Momentum-Averaged MA
balanced models, the total temperature Entropy-Averaged EA
ratio across all the interfaces is almost Flux Balanced MA
strictly maintained, demonstrating the (FBMA)
Flux Balanced
superior performance of flux balanced Flux
mixing models on flow conservation. Balanced EA (FBEA)
Experiment:
TA
Time-Averaged
Experiment:
MP
Total Pressure Ratio

Table 7.4.1 Nomenclature for different Mixing Models


Used in Study

ZHU YaLu, LUO JiaQi & LIU Feng, “Flow computations of multi-stages by URANS and flux balanced mixing
136

models”, Science China, Technological Sciences, July 2018 Vol.61 No.7: 1081–1091.
155

Compared with other flow variables, the relative 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 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 7.4.8 shows the span-wise distributions of total
pressure ratio and total temperature ratio on the first interface. The span-wise distributions of total

Figure 7.4.6 Contour of Relative Mach Number and Iso-Surface of Axial Velocity of Modified Rotor 67

pressure ratio by the mixing models perform rule-less variations, whereas the deviations from time-
averaged 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

Figure 7.4.7 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.)
156

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
7.4.6. The position of shock wave on the interface can also be clearly displayed by the contour of
static temperature in Figure 7.4.7. The shock wave injects onto the interface from 20% to 75%
spans, where the span-wise distributions of flow variables by different computation methods are not
consistent with each other as shown. 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 7.4.7. 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.
7.4.4 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.] 137. 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
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.
7.4.4.1 Transonic Turbine Stage Meshing and Flow Details

137Carlos 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.
157

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 Figure 7.4.8 Mesh for Transonic Turbine Stage - Upper Image
detect any flow convergence Depicted the Mesh at the Hub Surface while the Lower Image
problems, Represented Mesh used for the Blade Span
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 7.4.8.
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 stage inlet
we have imposed stagnation
pressure and temperature with
the corresponding flow angles
and at stage outlet static
Figure 7.4.9 Results of the Velocity Contours for a Radial Section
pressure is imposed. At the at Stator Mid Span using the Mixing Plane Approach
mixing plane interface also
158

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 7.4.9 presents the numerical results of velocity obtained for a
section at stator mid span. Although, numerical result shows a good agreement with the experimental
data, however, 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.]138 .
7.4.5 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]139, 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

Figure 7.4.10 Schematic View of Pitch-Wise Mixing Model

138 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.
139 Du Pengcheng a, Ning Fangfei, “Validation of a novel mixing-plane method for multistage turbomachinery

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


159

7.4.10-(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 7.4.10-(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 7.4.10-(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. 7.4.3
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.
7.4.5.1 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.
7.4.5.2 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 7.4.11-(A)). Meanwhile, the thick rotor wakes propagate through the second
mixing plane (MP2) and interact with the downstream stator (Figure 7.4.11-(B)). Thus, this test

(A) Static
(B) Entropy
Pressure

Figure 7.4.11 Instantaneous Distributions at 90% Span.


160

case 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 .
7.4.6 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,

PS = K R PR
Eq. 7.4.4
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.]140 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 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 Figure 7.4.12 Total Pressure Calculated by the Frozen Rotor
exchange strategies can be found

V. Brost, A. Ruprecht, and M. Maih, “Rotor-Stator interactions in an axial turbine, a comparison of transient
140

and steady state frozen rotor simulations”, Conference on Case Studies in Hydraulic Systems-CSHS03, 2003.
161

in141. (See Figure 7.4.12).

7.5 Un-Steady Treatment of Interface


7.5.1 Sliding Mesh (MRF)
Full unsteady simulations that integrate the governing equations in time can be performed to model
the nonlinear unsteady 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 row142. 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 Rai143.
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]144. 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 simulation145,
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 framework146. 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 7.5.1 A). The
halo cell approach: a halo-cell is created as a specular image of the interface cell (Figure 7.5.1 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 grids147.

141 NUMECA International, Brussels, “Fine/Turbo User Manual V8 (including Euranus)”, October 2007.
142 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.
143 M. Rai, “Application of domain decomposition methods to turbomachinery flows”, ASME Advances and

Applications in Computational Fluid Dynamics, volume 66, 1988.


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

Applications in Computational Fluid Dynamics, volume 66, 1988.


145 Reza Amini, “Using Sliding Meshes”.
146 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.
147 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.
162

Figure 7.5.1 Half Stencil and Full Stencil Reconstruction with: A) Intersection, B) Halo-Cell

7.5.2 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 He148 following the framework of Giles149 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
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 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

148 L. He, “Modelling issues for computation on unsteady turbomachinery flows. In Unsteady Flows in
Turbomachines”, Von K´arm´an Institute for Fluid Dynamics, 1996.
149 M. B. Giles, “An approach for multi-stage calculations incorporating unsteadiness”, ASME-GT92, number 282,

Cologne, Germany, 1992.


163

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
harmonics150. Figure
7.5.2 displays
Harmonic function
method in obtaining
relative velocities
(courtesy of
NUMECA.com). Since Figure 7.5.2 Relative Velocities Obtained using HB Techniques
the Harmonic method
is widely used, it is warranted a bit more exploring which will be dealt in the coming section.
7.5.3 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
unequal pitch problem between the blade
passages of neighboring rows. This
(ANSYS151, Galpin152), 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
maintaining full conservation. Multiple Figure 7.5.3 Phase shifted Periodic Boundary
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

150 S. Vilmin, E. Lorrain, and Ch. Hirsch, ” Unsteady flow modeling across the rotor/stator interface using the
non-linear harmonic method”, In ASME-GT06, number 90210, Spain, 2006.
151 ANSYS CFX Version 12 documentation, ANSYS Inc., 2009.
152 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.
164

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
boundaries153 (Figure 7.5.3).
7.5.4 Time Transformation Method (TT) using Phase-Shifted Periodic Boundary Conditions154
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 7.5.4. The Time Transformation method handles the

Figure 7.5.4 Phase Shifted Periodic Boundary Conditions

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 7.5.4. 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 )
U S1 (r, θ, z, t ) = U S2 (r, θ + PS , z, t − Δt ) Eq. 7.5.1

 U S1 (r, θ, z, t ) = U S2 (r, θ + PS , z, t )
Using the following set of space-time transformations to the problem above as:

Δt
r = r , θ = θ , z = z, t = t − Eq. 7.5.2
PR − PS

153 “A comparison of advanced numerical techniques to model transient flow in turbomachinery blade rows”,
Proceedings of ASME Turbo Expo 2011 GT2011.
154 ANSYS CFX-Solver Theory Guide, Release 15, 2013.
165

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 ΔtS
= Eq. 7.5.3
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. 7.5.4
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
instability.
7.5.5 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]155, and [Ning]156 developed a harmonic method in which the unsteady harmonics
are treated as perturbations. [Hall, Thomas, and Clark]157 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]158 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]159. In all these methods, the harmonic balance equations
are solved by introducing a pseudo-time derivative term and then marching the coupled equations

155 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.
156 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.


157 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.


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

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


159 He, L., 2010. “Fourier methods for turbomachinery applications”. Progress in Aerospace Sciences.
166

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.
7.5.5.1 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) , U (x + G , t) = U (x , t + Δt)
Eq. 7.5.5
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.
7.5.5.2 Boundary Conditions
We first consider the flow field kinematics of two adjacent blade rows where the first row has B1
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

N m1,m2 = m1B1 + m2 B2 Eq. 7.5.6

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 = m2 B2 (ω1 − ω2 ) , ω2,m1 = m1B1 (ω2 − ω1 ) Eq. 7.5.7

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 = m 2 2π , σ m1 ,2 = m1 2π Eq. 7.5.8
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:
167

M
U (x, t) =  Uˆ
m=− M
m (x) e imt
N −1 Eq. 7.5.9
where ˆ ( x) = 1  U
U
~ -imt n
m n ( x, t n ) e
N n =0

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. 7.5.10
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
̂ =𝐄𝐔
𝐔 ̃ or ̃ = 𝐄 −𝟏 𝐔
𝐔 ̂
Eq. 7.5.11
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

̃ = [𝐔
𝐔 ̃ 0 ,𝐔
̃1 ,𝐔
̃ 2 , ....... 𝐔
̃ N−1 ]T
Eq. 7.5.12
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. 7.5.13
Where the flux and source vectors F͂͂,͂͂ G̃ , and S͂͂͂͂ are evaluated using the corresponding time level
solution. The time derivative in Eq. 7.5.13 is evaluated by differentiating Eq. 7.5.11 with respect
to time as follows:
̃ ∂𝐄 −1
∂𝐔 ∂𝐄 −1
= ̂
𝐔= ̃ = [𝐃]𝐔
𝐔 ̃
∂t ∂t ∂t
Eq. 7.5.14
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. 7.5.15
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
168

derivative of primitive quantities, ∂Q/∂τ, with Q = {p, u, T}, is introduced into Eq. 7.5.15 to facilitate
solution of the steady harmonic balance equations by means of a time marching procedure. An Euler
implicit discretization in pseudo-time160 produces the following linearized system of equations:

𝛛𝐔 𝛛𝐒 𝛛𝐔
[ + Δτ ([𝐀] − + [𝐃] )] Δ𝐐′ = −Δτ 𝐑′
𝛛𝐐 𝛛𝐐 𝛛𝐐
Eq. 7.5.16
where R' is the discrete residual, 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 Eq.
7.5.16 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:

 U  S  ~
 + Δτ [ A] −  ΔQ = −Δτ R
 Q  Q 
Eq. 7.5.17
 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.
7.5.5.3 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 flows161. 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
7.5.5 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

160 Weiss, 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
161 L. He, "An Euler Solution for Unsteady Flows around Oscillating Blades", ASME, Journal of Turbomachinery,

Vol.112, No.4, pp.714-722, 1990.


169

passage under an inlet distortion of long circumferential wave length. Given only 3-5 harmonics were

Figure 7.5.5 Stagnation Pressure Contours under inlet distortion for NASA Rotor 67

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.
7.5.5.4 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
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 Figure 7.5.6 Instantaneous Pressure Distribution Within
the Compressor Stage Using (NLHB)
retained for the blade passing
170

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 7.5.6 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 and below are phase-shifted reconstructions included for clarity.
7.5.5.5 Case Study 2 - 3D Flow in Turbine Cascade
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 effects162.
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 7.5.7.
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
Figure 7.5.7 Computational Mesh for HB and TRS Methods
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 7.5.8 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.

162 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’, Russia.
171

TRS HB

Figure 7.5.8 Instantaneous Predictions of Turbulent Viscosity at Mid-Span for HB and TRS Solutions

7.5.6 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
at affordable computational cost [Rubino et al.]163. 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.]164.

163 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.
164 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.


172

7.5.6.1 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,
Germany165. 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 7.5.1. 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 7.5.1 Axial turbine simulation parameters
accuracy is achieved by MUSCL
reconstruction. For both unsteady
and steady simulations, non-reflective boundary conditions are imposed166 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) method167. 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. 7.5.18
where ⟨Ss;in⟩ and ⟨Ss;out⟩ are the stator inlet and outlet entropy values averaged over the boundary
using a mixed-out procedure168, 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.
7.5.6.2 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 7.5.9 shows the evolution in time of the
entropy generation, Sgen, calculated for a different number of time instances as well as the value given

165 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.
166 Giles, M. B.. “Nonreflecting boundary conditions for euler equation calculations”. AIAA journal, 1990.
167 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.


168 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.
173

by the steady state mixing-plane (MP) method.

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

Figure 7.5.9 Non-Dimensional Entropy Generation Using Unsteady (HB) vs Steady (MP)

7.5.6.2.1 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 7.5.1. It was also revealed at [Rubino et al.]169 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 7.5.9 (a), and about 3% in the case
of the stage total-to-static efficiency.
7.5.6.2.2 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 7.5.1). 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-
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 7.5.10 reports the non-dimensional
pressure contour plot, for both MP and HB simulations. Further details can be obtained from170.
7.5.6.3 Design Sensitivity
As mentioned in the CASE STUDY section, the entropy generation Sgen(U,a) defined by Eq. 7.5.18 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

169 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.
170 See Previous.
174

shown in [Rubino et al.]171 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. 7.5.19
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 obtained172. The AD and FD gradients, calculated with respect to a representative ensemble of 24

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

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

Figure 7.5.10 Case2: Non-Dimensional Pressure Contours for the Mixing Plane (MP) simulation
(a) , and Harmonic Balance (HB) at Different Time Instances (bcd)

171 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.
172 See Previous.
175

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 stability173. 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.
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 7.5.1). 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. 7.5.20
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.

173Gopinath, 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.
176

7.5.6.3.1 Case1 - Subsonic Stage


Figure 7.5.11 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 7.5.11(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:
• when compared to the sliding-
mesh interpolation used for (a) Case 1 - Stator
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 7.5.11 (b) the zone in the (b) Case 1 - Rotor
proximity of the rotor leading edge is
the one associated with the highest Figure 7.5.11 Mixing Plane vs Harmonic Balance
dsgen value. Such difference is mainly Normalized Entropy Generation Gradients Obtained with the
due to the simplification introduced Adjoint Solution
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.
7.5.6.3.2 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 cascade174. 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

174A. 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.
177

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.
7.5.6.4 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.

7.6 Case Study - Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design
Perspective175
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.

175Paolo Gaetani, “Stator-Rotor Interaction in Axial Turbine: Flow Physics and Design Perspective”, Chapter 5,
http://dx.doi.org/10.5772/intechopen.76009.
178

7.6.1 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
loading176. 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
budget177. 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
boundary layer properties may cause some pressure fluctuation on the cascade loading, as discussed
in178.
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 others179-180).
All these issues have been addressed both experimentally and by proper CFD simulations;
experiments require high promptness instrumentation like FRAPP (among others181) or LDV and PIV.

176 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.
177 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.


178 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.


179 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


180 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
181 Persico G, Gaetani P, Guardone A. Design and analysis of new concept fast-response pressure probes.

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


179

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
conditions182.
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)
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.
7.6.2 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

Ong J, Miller RJ. Hot streak and vane coolant migration in a downstream rotor. Journal of Turbomachinery. 7
182

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


180

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.
7.6.2.1 Stator Wake-Rotor Blade
Interaction Figure 7.6.1 Velocity triangles for the free stream
The stator wake can be regarded either (subscript FS) and the wake (subscript W) lows. V =
as a velocity defect or a loss filament. absolute velocity, W = relative velocity, U = peripheral
According to the first approach, the velocity
velocity triangle composition shows a
very different direction and magnitude for the relative velocity. Figure 7.6.1 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 bowed and
then chopped by the
Figure 7.6.2 Pattern of entropy evolution (bowing, chopping and transport) of
rotor leading edge.
the stator wake in the rotor channel, as foreseen by CFD
Later on, it is
181

transported inside the rotor channel, being smeared and showing two separate 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 7.6.2 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 different time even 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 183-184. 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.
7.6.2.2 Stator Secondary Lows-Rotor Blade Interaction
The basic mechanism for this kind of interaction is the same of the wake-blade one, the vorticial
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 7.6.3 and discussed in185. 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.

183 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
184 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.


185 Langston 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
182

Figure 7.6.3 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
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]186. 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.
Possible incidence angles to the stator additionally modulate the secondary lows. Positive incidence

186Perdichizzi A. Mach number effects on secondary low development of a turbine cascade. Journal of
Turbomachinery. 1990;112:643-651. DOI: 10.1115/1.2927705
183

angle strengthens secondary lows, as well as lower solidities, as a consequence of the higher blade
loading. Among others, a global review.
The low entering the rotor is then highly three dimensional and complex, as depicted in Figure 7.6.4.
In the case presented in Figure 7.6.4, 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

Figure 7.6.4 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).

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 7.6.5. 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 7.6.5(a)). The condition of constant inlet total
pressure both in the absolute and in the relative frame is, so far, an unrealistic condition; Figure
7.6.5( b) shows the periodic nonuniformities throughout the relative total pressure coefficient
(CPT,R).
184

Figure 7.6.5 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).

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.
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 7.6.6).
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.
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 Figure 7.6.6 Schematics of the stator vortical
one, tend to weaken it (and the same occurs for structure transport in the rotor passage
the hub ones). On the contrary, the stator shed
185

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 the rotor field.
Before discussing in detail the different time frames, it is straightforward to consider first the mean
flow (Figure 7.6.7 (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 reported in
Figure 7.6.7 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.

Figure 7.6.7 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).

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 7.6.8 how the 25 channels of the turbine rotor experience different low
conditions, each of them different with respect to the time mean one.
186

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 7.6.8 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
contrary have a clear periodic evolution with some instant where they do not exist.
187

Figure 7.6.9 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 7.6.7 (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 7.6.9 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).

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
188

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 7.6.10. With reference to the relative total pressure coefficient, the wider fluctuations are,
as qualitatively expected by Figure 7.6.8 and Figure 7.6.9 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 7.6.10 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 7.6.10 Standard deviation of the Cptr and δ for the different time frames. Experiments at the Fluid
machinery Lab. at Politecnico di Milano (Italy).

7.6.2.3 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 7.6.11, shows the relative total pressure coefficient and deviation angle standard deviations,
calculated among the different time instants, for a negative incidence conditions (Figure 7.6.11 a,
incidence at midspan = −10°) and a positive one (Figure 7.6.11 b, incidence at midspan = +10°).
189

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 behavior187.

Figure 7.6.11 Rotor loading effects on the stator-rotor interaction. Experiments at the Fluid machinery
Lab. at Politecnico di Milano (Italy).

7.6.2.4 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
suction side shock is stronger than the pressure side one. The suction side shock propagates

187Knobloch 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.
190

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
7.6.12; 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 7.6.12) and then reached the leading edge
(measuring point n° 2).

Figure 7.6.12 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.
191

7.6.3 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.
7.6.3.1 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 in188, 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 7.6.13. 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 7.6.14.

Figure 7.6.13 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).

188Gaetani 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
192

Overall, the stage experiences the


maximum efficiency for the design case
(Figure 7.6.15), 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 7.6.14 Rotor incidence fluctuations in
circumferential direction for the different axial gaps along
7.6.3.2 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 7.6.15 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
193

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 7.6.4 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.
7.6.3.3 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 from189, 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.]190, 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.

189 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
190 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
194

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.
195

8 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.

8.1 Centrifugal Compressor


Centrifugal compressors, as depicted in
Figure 8.1.1, sometimes termed Radial
compressors, are a sub-class of dynamic
axisymmetric work-absorbing
turbomachinery . The idealized compressive
191

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.
8.1.1 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
the flow continues to pass into and through the
centrifugal impeller, the impeller forces the flow Figure 8.1.1 Centrifugal impeller with a highly
to spin faster and faster. According to a form of polished surface likely to improve performance
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.
8.1.2 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.192,193 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
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

191 Shepard, Dennis G. “Principles of Turbomachinery”, McMillan. ISBN 0-471-85546-4. LCCN 56002849, 1956.
192 Lakshminarayana, B. (1996). “Fluid Dynamics and Heat Transfer of Turbomachinery”, New York: John Wiley
& Sons Inc. ISBN 0-471-85546-4.
193 Japikse, David & Baines, Nicholas C., “Introduction to Turbomachinery”, Oxford: Oxford University press.

ISBN 0-933283-10-5, 1997.


196

"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.
8.1.3 Components of a simple Centrifugal Compressor
A simple centrifugal compressor has four components: inlet, impeller/rotor, diffuser, and collector.
Figure 8.1.2 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
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)

Figure 8.1.2 Cut-Away View of a Turbocharger showing the Centrifugal Compressor

and both pressure and temperature instrumentation. All of these additional devices have important
uses in the control of the centrifugal compressor.
8.1.3.1 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.
8.1.3.2 Centrifugal Impeller
The key component that makes a compressor centrifugal is the centrifugal impeller, Figure 8.1.1
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
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
197

splitters" (every other inducer removed) and "w/o splitters" (all full blades). Figure 8.1.2-Figure
8.1.3 showing an open impellers with splitters. Most modern high efficiency impellers use "back
sweep" in the blade shape194-195. Euler’s pump and turbine equation plays an important role in
understanding impeller performance.
8.1.3.3 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 8.1.3 Jet Engine Cutaway Showing the Centrifugal Compressor among others

8.1.3.4 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
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.

194Japikse, David. “Centrifugal Compressor Design and Performance”. Concepts ETI . ISBN 0-933283-03-2.
195 Aungier, Ronald H. (2000). “Centrifugal Compressors, A Strategy for Aerodynamic Design and Analysis”. ASME
Press. ISBN 0-7918-0093-8.
198

8.1.4 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.
8.1.4.1 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.
8.1.4.2 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.
8.1.4.3 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.
8.1.4.4 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.
8.1.4.5 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.
8.1.4.6 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
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.
8.1.4.7 Air Separation Plants to Manufacture Purified End Product Gases
199

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.
8.1.4.8 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.

8.2 Radial Turbine


A radial turbine is a turbine in which the flow of the working fluid is radial to the shaft196. 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.
8.2.1 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.

Figure 8.2.1 Ninety Degree Inward-Flow Radial Turbine Stage

8.2.2 Types of Radial Turbines


Radial flow turbines may be classified as:

196 From Wikipedia, the free encyclopedia.


200

• Inward flow radial (IFR) turbines


➢ Cantilever turbine
➢ 90 degree turbine (see Figure 8.2.1)
• Outward flow radial (OFR) turbines (see Figure 8.2.2)
8.2.2.1 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.
8.2.2.2 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 8.2.2 Outward Flow Radial Turbine

8.2.2.3 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
8.2.2).
201
202

9 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 components197. 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.

9.1 Quasi-3D (Q3D) or 3D Simulation


9.1.1 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.
9.1.2 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.
9.1.3 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
9.1.1 shows the flow range from 2D to 3D by different vendor codes.

197 CFD on line series.


203

3D transient Solver from ANSYS

Q3D Solver from


NUMECA
2-D Steady state
transonic viscous flow

Figure 9.1.1 Different Flow (2D, Q3D, and full 3D)

9.2 Single vs Multi-Stage Analysis


9.2.1 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
204

inherently unsteady and involves the full-annulus flow field. Also, at part speeds, the predicted
efficiency values are noticeably lower than the measured values.
9.2.2 Multi-Stage Analysis
9.2.2.1 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.
9.2.2.2 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.
9.2.2.3 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.
9.2.2.4 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 9.2.2 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
205

Figure 9.2.2 Full Blade Simulation using Harmonic Balanced Method (Courtesy of CD-adapco)

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
formulations. The nonlinear harmonic approach have been extended to effectively solve rotor-
rotor/stator-stator interactions in multistage turbomachines 198.
9.2.2.5 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.
9.2.2.6 Other Advanced Multi-Stage Methods
Time-inclined, Adamczyk stresses, etc.

9.3 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.

198 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.
206

9.4 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
(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 authors199 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. 200
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]201 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 effect202 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 203. 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 neglected204. 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;

199 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).
200 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.
201 Pullan, G. (2006). “Secondary flows and loss caused by blade row interaction in a turbine stage”. ASME Journal

of Turbomachinery, 128(3), 484–491.


202 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.


203 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.


204 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.
207

• Rotating stall: typical of the compressors, is caused by the blockage of some vanes due to the
wrong incidence which causes flow separation;
• 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 9.4.1 Transient Blade Row Extensions Enable Efficient Multi-
Stage CFD Simulation (Courtesy of ANSYS.com)
simulation and average the
results if you don't obtain a
converged steady solution. Figure 9.4.1 shows the transient blade row extensions enable efficient
multi-stage CFD simulation (courtesy’s of ANAYS.com).

9.5 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
208

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 9.5.1 shows
a typical meshing for a 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
Figure 9.5.1 Typical Meshing of a Turbomachinery
approach with special attention to Stage
steady/unsteady analysis and to
turbulence modelling (Montomoli et al.,
2015)205. At the topological level of a
blade, it is customary to use a H-O-H
meshing block, with an O block around
the blade and H blocks placed in both the
upstream and downstream of the O block
(Kato et al., 2010) and see (Figure 9.5.2).
9.5.1 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
coarse mesh is often able to obtain a good
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 Figure 9.5.2 Computational Block for a Single Blade
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.

205F. 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.
209

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 good wall-function mesh has 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 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 9.5.3
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 Figure 9.5.3 Multi-Block Grid for the Space Shuttle Main Engine
the trailing edge region is too Fuel Turbine
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 resolution206.
9.5.2 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]207 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

206 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.
207 Sergiy Yershov, Viktor Yakovlev, “Mesh Resolution Effect On 3d Rans Turbomachinery Flow Simulations”,

Institute For Mechanical Engineering Problems Of Nasu, Kharkiv, Ukraine, 2016.


210

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.
9.5.2.1 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:
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.
9.5.2.2 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]208.

208Sergiy Yershov, Viktor Yakovlev, “Mesh Resolution Effect On 3d Rans Turbomachinery Flow Simulations”,
Institute For Mechanical Engineering Problems Of Nasu, Kharkiv, Ukraine, 2016.
211

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.
9.5.3 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
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.
9.5.4 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.
212

9.6 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 interfaces209. 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
rows. [Giles]210 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 Figure 9.6.1 Pressure contour plot, 2nd order spatial
which the boundaries are often discretization scheme
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

209 Rodrick V. Chima, “Calculation of Multistage Turbomachinery Using Steady Characteristic Boundary
Conditions”, AIAA 98-0968.
210 Giles, Michael B., “Nonreflecting Boundary Conditions for Euler Equation Calculations,” AIAA Journal, Vol. 28,

No. 12, Dec. 1990, pp. 2050-2058.


213

enhanced due to an improvement of the transmission of outgoing waves, allowing smaller meshes to
be used211. Figure 9.6.1 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 result of the reflectivity of the 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 comparison212.
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 al213 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 parameter214.

9.7 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

211 “Three-Dimensional Nonreflecting Boundary Conditions for Swirling Flow in Turbomachinery”, Journal of
Propulsion and Power Vol. 23, No. 5, September–October 2007.
212 F. De Raedt, “Non-Reflecting Boundary Conditions for non-ideal compressible fluid flows”, Master of Science at

the Delft University of Technology, 2015.


213 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).
214 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.


214

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.

9.8 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.
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 effects215.
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.
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 9.8.1 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.

215 ANSYS Blog.


215

Figure 9.8.1 Analysis provided vibration required for flutter analysis

9.8.1 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 9.8.2 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 9.8.2. 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 9.8.2. These are
commonly called nodal diameters. Hence the natural frequency and nodal diameter are required to
describe a bladed disk mode216.

9.9 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

216Mohamed Hassan, “Vibratory Analysis of Turbomachinery Blades”, Master of Mechanical Engineering,


Rensselaer Polytechnic Institute, Hartford, Connecticut, December, 2008.
216

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.

9.10 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
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 217 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.

9.11 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.

9.12 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-

217 Latin phrase meaning “for this purpose”.


217

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.

9.13 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.

9.14 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
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.
9.14.1 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 blades218. 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.

218A.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.
218

9.15 Literature Review and Parallel Processing Tools


A review of CFD analysis for turbo-machineries is acquired by (Pinto et al., 2016)219. 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.
• 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 results (Montomoli et al., 2015) 220. 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 United Device Architecture)
3. PVM (Parallel Virtual Machine)
4. Hybrid OpenMP + MPI
5. Tri-level hybrid OpenMP + MPI+CUDA
6. OpenGL (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

219 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.
220 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.
219

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.
A tabulated results of
literature reviews of
available commercial codes
involving polarizations can
be obtained by (Pinto, et al.,
2017)221, with results
represented in Table
9.15.1.

9.16 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 Table 9.15.1 Parallel computing tools adopted for parallelization of
approach, there will always CFD code used for turbomachinery – Courtesy of Pinto et al.
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
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

221 Pinto, R.N., Afzal, A., D’Souza, L.V.


et al. Computational Fluid Dynamics in Turbomachinery: A Review of State
of the Art. Arch Computational Methods Eng 24, 467–479 (2017) https://doi.org/10.1007/s11831-016-9175-
2
220

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 turbomachinery222.

222F. 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