5%? ?.:.n, A) m '1 ~ ‘73:: :1 .7} e . ‘h‘ I 4?; This is to certify that the dissertation entitled PERFORMANCE PREDICTION AND PRELIMINARY DESIGN OF WAVE ROTORS ENHANCING GAS TURBINE CYCLES presented by PEZHMAN AKBARI has been accepted towards fulfillment of the requirements for the Ph.D. degree in Mechanical Engineering A I ,' .1" ;: .‘ L /{ j: 1’ Ma‘fir‘Profe'ssor’s Signature 8/23/2004 Date MSU is an Affirmative Action/Equal Opportunity Institution .—.-.—-—..—--... ---. ms.-o-n-o-o-n-a—c-o—o-o-I-n-n-u-o-o—o—‘-o-n— on i- .— LIBRARY Michigan State University PLACE IN REIURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested DATE DUE DATE DUE DATE DUE é 2/05 a/cmcromom.mms PERFORMANCE PREDICTION AND PRELIMINARY DESIGN OF WAVE ROTORS ENHANCING GAS TURBINE CYCLES By Pezhman Akbari A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2004 ABSTRACT PERFORMANCE PREDICTION AND PRELIMINARY DESIGN OF WAVE ROTORS ENHANCING GAS TURBINE CYCLES By Pezhman Akbari Wave rotors as a family of unsteady-flow devices have shown unique capabilities to enhance the performance and operating characteristics of a variety of engines and machinery utilizing thermodynamic cycles. The wave rotor is an unsteady-flow machine that utilizes compression and expansion waves to exchange energy between fluids with different pressures. The first part of this study presents several thermodynamic cycle analyses proving the performance improvement of small gas turbines (microturbines) by implementing various advantageous four-port wave rotor topping cycles. General performance maps are generated showing the design space and optima for baseline and topped engines. In the second part of this work, a one-dimensional analytical gasdynamic model of the hi gh-pressure phase is suggested to calculate flow characteristics inside the wave rotor channels. Useful design parameters such as port widths and rotor size are determined by computing transit times of the waves traveling inside the channels. Reasonable agreement has been found comparing the predicted results of the analytical design procedure with the available numerical data. Using the commercial software FLUENT, some CFD simulations are performed showing that commonly available CFD software can be utilized for wave rotor simulations and to support their design. Overall, the predicted results are in good agreement with established wave rotor theories. Finally, several innovative wave rotor concepts and designs including have been studied. To My Family iii ACKNOWLEDGMENTS I would like first to express my gratitude to my supervisor Prof. Norbert Miiller for providing an opportunity to make the present research a true learning process. He generously allowed me to start this work in order to learn and satisfy my own curiosity. His sense of guidance, unending patience, refreshing humor, and friendly suggestions have made my research experience both fruitful and enjoyable. His encouragement and timely advice have always helped me to overcome the difficulties I faced during my research. I consider myself lucky to have Prof. Norbert Miiller as my supervisor. I have had a privilege to learn a lot from my Ph.D. committee members. As a former student of Prof. Indrek Wichman, I admire his inspiring work and his continuous support. Prof. Abraham Engeda deserves my special thanks as one of the most supportive advisors through my research, offering very helpful discussions and suggestions. I am also grateful to Prof. Farhad Jaberi on my Ph.D. committee for his valuable suggestions. He has always been a valuable mentor and an excellent model for me. Special thanks to Prof. Charles MacCluer for serving on my Ph.D. committee. His helpful advice and feedback have been very beneficial for me. I am also very thankful to Prof. Razi Nalim at the Purdue School of Engineering and Technology at IUPUI. As an acting member of my Ph.D. committee and top specialist in his field, he has provided me with continuous and valuable guidance in improving this work. I would particularly like to thank Prof. Jenasz Piechna at Warsaw University of Technology in Poland with whom I have had many constructive discussions. I appreciate his unique talents and brightness and enjoy his friendship. iv Thanks to Prof. Manoochehr Koochesfahani for assisting me to begin my Ph.D. at MSU. I am honored to had a chance to work in his laboratory in the early stages of my Ph.D. program. I am also very grateful to my former supervisor Prof. Kaveh Ghorbanian at Sharif University of Technology in Iran who first inspired me to work on such a challenging topic. Working with him was a valuable and enjoyable undertaking. This work would have never been completed without the support of several colleagues in the MSU Turbomachinery Laboratory. I am thankful for the collaboration and friendship of Florin Iancu whose assistance in numerical simulations performed in this work is appreciated. His patience during our collaboration is appreciated. I would like to thank Amir A. Kharazi for a productive collaboration, leading to our several technical publications. The materials presented in Chapter 6 are the result of my collaboration with the above friends at the MSU wave rotor team. I am also very thankful to Azadeh Behinfar as a real friend and a great source of support. Several of the 3D sketches presented here are results of her excellent skills. Thanks for keeping me happy during those difficult moments. Finally and most importantly, I certainly remain indebted to my parents and all other family members for their continuing love, encouragement, and endless support through this journey. I could never pay them back what they have given me. TABLE OF CONTENTS LIST OF TABLES VIII LIST OF FIGURES IX CHAPTER 1: INTRODUCTION 1 1.1 Unsteady-Flow Devices ......................................................................................................... l 1.2 Wave Rotors .......................................................................................................................... 4 1.3 Wave Rotor Applications ...................................................................................................... 6 1.3.1 Gas Turbine Applications ............................................................................................... 7 CHAPTER 2: HISTORICAL REVIEW OF WAVE ROTOR TECHNOLOGY ..... 14 CHAPTER 3: WAVE ROTOR THEORY 48 3.1 Energy Exchange with Waves ............................................................................................. 48 3.2 Charging Process in Wave Rotors ....................................................................................... 53 3.3 Principles of F our-Port Wave Rotor Operation ................................................................... 57 3.3.1 Through-Flow versus Reverse-Flow Wave Rotors ...................................................... 57 3.3.2 How Does it Work Inside? ........................................................................................... 59 3.4 Gasdynamic Equations ........................................................................................................ 63 3.4.1 Moving Normal Shock Wave Relations ....................................................................... 63 3.4.2 Expansion Wave Relations ........................................................................................... 66 CHAPTER 4: THERMODYNAMIC ANALYSIS 70 4.1 Microturbines ...................................................................................................................... 70 4.2 Gas Turbine without Recuperation ...................................................................................... 72 4.2.1 Implementation Cases .................................................................................................. 72 4.2.2 Thermodynamic Calculations ...................................................................................... 76 4.2.3 Predicted Performance Results ..................................................................................... 86 4.2.4 Comparison Between Adding a Second Compressor Stage with Wave-Rotor-Topping ............................................................................................................................................. 105 4.2.5 Substituting the Compressor in the C-60 Engine with the C-30 Compressor plus Wave Rotor .................................................................................................................................... 111 4.2.6 Effect of Compressor Inlet Temperature .................................................................... 112 4.3 Gas Turbine with Recuperation ......................................................................................... 117 4.3.1 Thermodynamic Calculations .................................................................................... 119 4.3.2 Predicted Performance Results ................................................................................... 122 4.4 Turbojet Engines Topped with Wave Rotors .................................................................... 128 4.4.1 Thermodynamic Calculations .................................................................................... 130 4.4.2 Predicted Performance Results ................................................................................... 134 CHAPTER 5: PRELIMINARY WAVE ROTOR DESIGN 138 5.1 Analytical Approach .......................................................................................................... 138 5.1.1 Charging Process with a Single Shock Wave ............................................................ 140 5.1.2 Charging Process with Two Shock Waves ................................................................. 146 5.1.3 Validation ................................................................................................................... 163 5.2 Numerical Simulation ........................................................................................................ 165 5.2.1 Approach .................................................................................................................... 167 5.2.2 Numerical Results ...................................................................................................... 172 vi CHAPTER 6: INNOVATIVE WAVE ROTOR DESIGNS AND APPLICATIONS 186 6.1 Radial Wave Rotor Concept .............................................................................................. 186 6.1.1 Radial-Flow Wave Rotor with Straight Channels ...................................................... 186 6.1.2 Radial-Flow Wave Rotor with Curved Channels ....................................................... 190 6.1.3 Radial-Flow Wave Rotor Concept ............................................................................. 193 6.2 Condensing Wave Rotor .................................................................................................... 197 6.2.1 How Does a Condensing Wave Rotor Work? ............................................................ 198 6.2.2 Performance Evaluation ............................................................................................. 204 6.3 Wave Rotors for Ultra—Micro Gas Turbines ...................................................................... 206 6.3.1 Performance Enhancement of UuGT Using Wave Rotors ......................................... 207 6.3.2 Conceptual Designs of Wave Rotor Implementation into UuGT .............................. 209 CONCLUSION 213 REFERENCES 217 APPENDICES 233 Appendix A: Shock Waves and Diffusers ............................................................................... 234 Appendix B: Temperature-Entropy Diagram Construction ..................................................... 239 vii LIST OF TABLES Table 1: Baseline engine data, assuming T 0 =300 K, Cpa,,=1.005 kJ/kgK, Cpgas=1.148 kJ/kgK, ya, =1.4,yga,=l.33 ...................................................................................................................... 77 Table 2: Performance comparison between baseline engines and five cases of wave-rotor-topping with a wave rotor pressure ratio of 1.8 ................................................................................ 101 Table 3: Performance comparison between adding a conventional second compressor stage and wave-rotor-topping Case A and E with a wave rotor pressure ratio of 1.8; baseline engine C- 30 ......................................................................................................................................... 108 Table 4: Performance comparison between adding a conventional second compressor stage and wave-rotor-topping Case A and E with a wave rotor pressure ratio of 1.8; baseline engine C- 60 ......................................................................................................................................... 109 Table 5: Ambient temperature effect on performance - comparison between baseline engines and Cases A and B of wave-rotor—topping with PRW =1.8; baseline engine C-3O ..................... 114 Table 6: Ambient temperature effect on performance - comparison between baseline engines and Cases A and B of wave-rotor-topping with PRW =1.8; baseline engine 060 ..................... 114 Table 7: Performance comparison between baseline engines and two cases of wave-rotor-topping with a wave rotor pressure ratio of 1.8 ................................................................................ 128 Table 8: Turbojet baseline engine data, assuming T 0 =300 K, Cpai,=1.005 kJ/kgK, Cpgas =1 . 148 kJ/kgK, yai,=1.4 , ygm=1.33 ................................................................................................. 130 Table 9: Performance comparison between baseline turbojet engine and five cases of wave-rotor- topping with a wave rotor pressure ratio of 1.8 ................................................................... 135 Table 10: Enhanced engine data, assuming PRW = pmpfl =1.8 for Case ................................... 143 Table 11: Flow properties inside rotor channels for L=20 cm ..................................................... 152 Table 12: Wave rotor design parameters for L=20 cm, N=30, nm,0,=10000 rpm, and m, =03 l 8 kg/s, ..................................................................................................................................... 152 Table 13: Inlet and outlet port data of University of Tokyo wave rotor, taken fi'om Ref. [215] 163 Table 14: Comparison between the analytical model and the numerical data of the Tokyo wave rotor ..................................................................................................................................... 165 viii LIST OF FIGURES Figure 1: Comparison of pressure gain (local pressure ratio) of moving shock and steady-flow isentropic diffuser for y=1.4 .................................................................................................... 2 Figure 2: Shock wave, diffuser, and compressor isentropic efficiencies as functions of pressure gain .......................................................................................................................................... 3 Figure 3: Schematic configuration of a typical wave rotor .............................................................. 4 Figure 4: Schematic configuration of a simple gas turbine engine .................................................. 7 Figure 5: Schematic temperature-entropy diagram for a gas turbine with low and high pressure ratios ........................................................................................................................................ 9 Figure 6: Schematic of a gas turbine topped by a four-port wave rotor ........................................ 10 Figure 7: Schematic example of the physical implementation of a wave rotor in a gas turbine (exploded view, piping not shown) ....................................................................................... 10 Figure 8: Schematic T-s diagrams for a gas turbine with and without a wave rotor ..................... 12 Figure 9: Wave rotor as a topping stage for the locomotive gas turbine, taken from Ref. [17].... 16 Figure 10: The Comprex®, taken from Ref. [34] .......................................................................... 17 Figure 11: Photograph of the CAL Wave Supercharger, taken from Ref. [5] ............................... 21 Figure 12: Schematic of a double wave rotor cycle, taken from Ref. [98] .................................... 22 Figure 13: The experimental divider test rig at Imperial College, taken from Ref. [5] ................. 23 Figure 14: The Pearson rotor (left) and rear and front stator plates (right), taken from Ref. [124] ............................................................................................................................................... 24 Figure 15: Ideal wave diagram of the Klapproth rotor, taken from Ref. [126] .............................. 28 Figure 16: Ideal wave diagram of the GPC rotor, taken from Ref. [109] ...................................... 30 Figure 17: Schematic of the MSNW experimental set up, taken from Ref. [21] ........................... 33 Figure 18: Conceptual design of a turbofan engine incorporating a wave rotor, taken from Ref. [138] ...................................................................................................................................... 35 Figure 19: Four-port wave rotor of NASA .................................................................................... 39 Figure 20: Temperature distribution of partition exit flow, taken from Ref. [204] ....................... 43 Figure 21: Rotary Wave Ejector Pulse Detonation Engine, taken from Ref. [210] ....................... 44 Figure 22: Wave Rotor PDE the ‘CVC’ Engine, taken from Ref. [211] ....................................... 45 Figure 23: University of Tokyo single-channel test rig, taken from Ref. [214] ............................ 46 Figure 24: Historical perspective of wave rotor technology. Red: gas turbine application, Green: IC engine supercharging, Blue: refrigeration cycle, Pink: pressure divider and equalizer, Purple: wave superheater, Orange: internal combustion wave rotors, Black: general applications ........................................................................................................................... 47 Figure 25: Flow in a shock tube after the diaphragm is ruptured .................................................. 49 Figure 26: Closed tube containing stationary gas .......................................................................... 49 Figure 27: Left end opening to gas at higher pressure ................................................................... 50 Figure 28: Lefi end opening to gas at lower pressure .................................................................... 51 Figure 29: Opened tube containing flowing gas ............................................................................ 52 Figure 30: Left end closing to generate an expansion wave .......................................................... 52 Figure 31: Right end closing to generate a shock wave ................................................................. 53 Figure 32: Wave diagram of charging process with two primary shock waves ............................ 54 Figure 33: Wave diagram of charging process using single shock wave and expansion wave 56 Figure 34: Schematic of a gas turbine topped by a reverse-flow four-port wave rotor ................. 58 Figure 35: Wave diagrams for through-flow (left) and reverse-flow (right) four-port wave rotors, taken form Ref. [168] ............................................................................................................ 61 Figure 36: Transformation from a moving shock wave to a stationary shock wave ..................... 64 Figure 37: Illustration of the characteristic lines in the x-t plane ................................................... 67 ix Figure 38: Lefi end closing to generate an expansion wave .......................................................... 67 Figure 39 : Schematic T-s diagrams for a baseline cycle and five wave-rotor-topped cycles ....... 73 Figure 40: Pressure gain ratio across the wave rotor versus temperature ratio across it ................ 83 Figure 41: Pressure gain ratio across the wave rotor versus wave rotor compression ratio .......... 84 Figure 42: T-s diagrams for the C-30 and C-60 wave-rotor—topped engines for Case A implementation ...................................................................................................................... 87 Figure 43: Thermal efficiency, specific work, and SF C for the wave-rotor-topped engines versus the wave rotor pressure ratio and overall pressure ratio, Case A consideration .................... 89 Figure 44: Relative values of thermal efficiency, specific work, and SF C for the wave-rotor- topped engines versus the wave rotor pressure ratio and overall pressure ratio, Case A consideration ......................................................................................................................... 90 Figure 45: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case B implementation ...................................................................................................................... 92 Figure 46: Thermal efficiency, specific work, and SF C for the wave-rotor-topped C-30 engine versus the wave rotor pressure ratio and compressor pressure ratio, Case B consideration 94 Figure 47: Relative values of thermal efficiency, specific work, and SF C for the wave-rotor- topped C-30 engine versus the wave rotor pressure ratio and compressor pressure ratio, Case B consideration ............................................................................................................. 94 Figure 48: Thermal efficiency, specific work, and SF C for the wave-rotor-topped C-30 engine versus the wave rotor pressure ratio and overall pressure ratio for Case A (solid) and versus wave rotor pressure and compressor pressure ratio for Case B (dashed) .............................. 97 Figure 49: Thermal efficiency, specific work, and SF C for the wave-rotor—topped C-6O engine versus the wave rotor pressure ratio and overall pressure ratio for Case A (solid) and versus wave rotor pressure and compressor pressure ratio for Case B (dashed) .............................. 97 Figure 50: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case C implementation ...................................................................................................................... 98 Figure 51: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case D implementation ...................................................................................................................... 99 Figure 52: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case E implementation .................................................................................................................... 100 Figure 53: Performance map for wave-rotor-topping of the C-30 engine, Cases A, B, and D... 102 Figure 54: Performance map for wave-rotor-topping of the C-60 engine, Cases A, B, and D... 102 Figure 55: Performance map for wave-rotor—topping of the C-30 engine, Cases C and E ......... 103 Figure 56: Performance map for wave-rotor-topping of the 060 engine, Cases C and E ......... 103 Figure 57: T-s diagrams for baseline cycle, conventional cycle with two-stage compressor (double compression work) and wave-rotor-topped cycle Case E with PRW =1.8, 030 engine .................................................................................................................................. 109 Figure 58: Thermal efficiency, specific work, and specific fuel consumption of the wave-rotor- topped C-60 engine using the C-30 compressor versus wave rotor compression ratio and compressor pressure ratio for an overall pressure ratio fixed at 4.8 (yellow points for original 060 baseline engine, orange points for wave-rotor-topped C—60 engine with C-3O compressor) ......................................................................................................................... 1 12 Figure 59: Effect of ambient temperature: absolute and relative changes of thermal efficiency, specific work, and SF C versus wave rotor pressure ratio for Case A ................................. 115 Figure 60: Effect of ambient temperature: absolute and relative changes of thermal efficiency, specific work, and SF C versus wave rotor pressure ratio for Case B ................................. 116 Figure 61: Schematic of a recuperated gas turbine topped by a four-port wave rotor ................. 117 Figure 62: T-s diagrams for recuperated baseline and wave-rotor-topped cycles, Case A .......... 118 Figure 63: T-s diagrams for recuperated baseline and wave-rotor-topped cycles, Case B .......... 118 Figure 64: Thermal efficiency, specific work, and SF C of the enhanced C-30 recuperated engine versus the wave rotor pressure ratio and overall pressure ratio for Case A ........................ 124 Figure 65: Relative values of thermal efficiency, specific work, and SF C of the enhanced C-30 recuperated engine versus the wave rotor pressure ratio and overall pressure ratio for Case B ............................................................................................................................................. 124 Figure 66: Thermal efficiency, specific work, and SF C of the enhanced C-3O recuperated engine versus the wave rotor pressure ratio and compressor pressure ratio for Case B ................. 126 Figure 67: Relative values of thermal efficiency, specific work, and SF C of the C-30 recuperated engine topped with a wave rotor versus the wave rotor pressure ratio and compressor pressure ratio for Case B ..................................................................................................... 126 Figure 68: Schematic of a turbojet engine topped by a four-port wave rotor .............................. 129 Figure 69 : Schematic T-s diagrams for a baseline turbojet cycle and five different wave-rotor- topped cycles ....................................................................................................................... 129 Figure 70: Performance maps for wave-rotor-topping of the C-30 turbojet engine, Cases A, B, and D ................................................................................................................................... 136 Figure 71: Performance map for wave-rotor—topping of the C-30 turbojet engine, Cases C and E ............................................................................................................................................. 137 Figure 72: Wave diagram of high-pressure part with only one shock wave ................................ 141 Figure 73: Wave diagram of high-pressure part with two primary shock waves ........................ 147 Figure 74: Preliminary port designs of the charging process, corresponding to the wave diagram sketched in Figure 73 .......................................................................................................... 153 Figure 75: Outlet port width comparison before and after rotor shortening ................................ 154 Figure 76: Variation of 175 and HR versus PRW for a constant value of 17comb ............................. 157 Figure 77: Shock strengths 175 , 17R , entropy production As/R , and exit velocity u; versus PR w and [Lamb ..................................................................................................................... 158 Figure 78: Variation of Lg“, /L versus PRW and 11%,, ................................................................. 161 Figure 79: Variation of dimensionless values of (WT- WWJ/ WT versus PRW and 17mm), ............ 162 Figure 80: University of Tokyo wave rotor, taken from Ref. [215]. ........................................... 164 Figure 81: Non-dimensional total temperature contour of the Tokyo wave rotor, taken from Ref. [215] .................................................................................................................................... 168 Figure 82: Non-dimensional total pressure contour of the Tokyo wave rotor, taken from Ref. [215] .................................................................................................................................... 169 Figure 83: Generated mesh .......................................................................................................... 171 Figure 84: Finer mesh at port interface and channels, inlet high-pressure port ........................... 171 Figure 85: Interface and gap between stationary ports and moving channels ............................. 172 Figure 86: Total pressure contours, the effect of high-pressure gas inlet port on the lst channel ............................................................................................................................................. 176 Figure 87: Total pressure contours, the effect of high-pressure gas inlet port on the lst and 2nd channels ............................................................................................................................... 177 Figure 88: Total pressure contours, traveling of compression waves toward end wall ............... 178 Figure 89: Total pressure contours, generation of reflected shock wave at two different time steps ............................................................................................................................................. 179 Figure 90: Total pressure contours, air scavenging process at two different time steps .............. 180 Figure 91: Total pressure contours, ingestion of fresh air at two different time steps ................. 181 Figure 92: Total pressure contour, end of the first operating cycle ............................................. 182 Figure 93: Axial component of relative velocity vector, gas inlet port and channels .................. 182 Figure 94: Axial component of relative velocity, lower and upper comers of gas inlet port ....... 183 Figure 95: Axial component of relative velocity, air outlet port .................................................. 184 Figure 96: Axial component of relative velocity, gas outlet port ................................................ 184 Figure 97: Axial component of relative velocity, air inlet port .................................................... 185 Figure 98: Mesh for multi cycles ................................................................................................. 185 Figure 99: Radial—flow wave rotor with straight channels ........................................................... 187 xi Figure 100: Reverse-flow wave disk with straight channels ....................................................... 188 Figure 101: Wave disc with straight channels: contour plots of local pressure (top), local temperature (middle), and velocity (bottom) at time=8.6e-O4 s .......................................... 189 Figure 102: Radial-flow wave rotor with curved channels .......................................................... 190 Figure 103: Reverse-flow wave disk with curved channels ......................................................... 191 Figure 104: Wave disc with curved channels: contour plots of local pressure (top), local temperature (middle), and velocity (bottom), at time=1.73e-3 s ......................................... 192 Figure 105: Stack of wave discs with straight channels .............................................................. 193 Figure 106: Stacked radial wave discs and radial compressor ..................................................... 194 Figure 107: Cut-view of a radial wave rotor topping a gas turbine ............................................. 195 Figure 108: Flow through an internal combustion radial wave rotor topping a gas turbine ........ 196 Figure 109: Component parts of a radial wave rotor topping a gas turbine ................................. 197 Figure 110: Schematic of a R718 cycle with direct condensation and evaporation .................... 199 Figure 111: Schematic of a R718 cycle enhanced by a three-port condensing wave rotor substituting for the condenser and for one compressor stage .............................................. 199 Figure 112: Schematic of a three-port condensing wave rotor .................................................... 200 Figure 113: Regions modeled for each compression and condensation ...................................... 200 Figure 114: Schematic wave and phase-change diagram for the three-port condensing wave rotor (high-pressure part) ............................................................................................................. 200 Figure 115: Schematic p-h diagram of a R718 baseline cycle and a wave rotor enhanced cycle ............................................................................................................................................. 203 Figure 116: Schematic T-s diagram of a R718 baseline cycle (cooling water cycle not shown) and a wave rotor enhanced cycle ............................................................................................... 203 Figure 117: Performance map representing maximum performance increase and optimum wave rotor pressure ratios ............................................................................................................. 205 Figure 118: Efficiency trend of compression process. Green: wave rotor efficiency, Red: compressor efficiency ......................................................................................................... 209 Figure 119: First conceptual design of an UuGT equipped with a four-port wave rotor ............ 210 Figure 120: Second conceptual design of an UuGT equipped with a four-port wave rotor ........ 211 Figure 121: Third conceptual design of an UuGT equipped with a four-port wave rotor ........... 212 xii CHAPTER 1: INTRODUCTION 1.1 Unsteady-F low Devices Oscillatory and pulsatile fluid motion has been neatly utilized in nature, yet is comparatively poorly studied by engineers despite the invention of cyclically operating engines and machines. The potential for utilizing unsteady flows has been recognized since the early twentieth century but neglected as long as substantive improvements could be made to conceptually simple steady-flow or semi-static devices. Also, the inherent non-linearity of large—amplitude wave phenomena in compressible fluids and unusual geometry of non-steady flow devices necessitates detailed calculations, which until recently were too laborious or expensive or imprecise. By understanding and exploiting complex unsteady flows, a quantum increase in engine performance is possible. Often it has been feasible to simplify the hardware of engines, making them less costly, more responsive, and more durable by employing unsteady-flow processes. Shock tubes, shock tunnels, pulse combustors, pulse detonation engines, and wave rotors are a few examples of unsteady-flow devices. The basic concept underlying these devices is the transfer of energy by pressure waves. By generating compression and expansion waves in appropriate geometries, wave machines can transfer energy directly between different fluids without using mechanical components such as pistons or vaned impellers. In fact, these devices properly represent applications of classical, unsteady, one-dimensional, compressible flow theory. The major benefit of these unsteady-flow machines is their potential to generate much greater pressure rises than those obtained in steady-state flow devices [1, 2]. For instance as illustrated in Figure 1, for the same change from a given inlet Mach number M, to a given outlet Mach number M2 , the local (static) pressure ratio p2 /p, across a single shock wave (solid line) moving in a frictionless channel is always much greater than that obtained by an isentropic deceleration in a 100% efficient diffuser (dotted line). However, it must be noted that friction effects always exist, and hence pressure gains are actually less than those predicted by Figure 1. 770,17 =100% M Mlfi— M2 E, E, 2.5 . (I) <2) 1) (2) Moving shock wave Diffuser p2/p1 N 1.5 - 0 0.2 0.4 0.6 0.8 1 M1 Figure 1: Comparison of pressure gain (local pressure ratio) of moving shock and steady- flow isentropic diffuser for y=1.4 It is also worthwhile to note that shock wave compression is a relatively efficient process as indicated in Figure 2, where shock isentropic efficiency ”Shock (red) is compared with compressor isentropic efficiency neompmm, (green) and diffuser isentropic efficiency 770mm (blue). Figure 2 shows variations of these parameters as functions of the pressure gain p; /p, obtained by a moving shock wave in a frictionless channel, by a compressor with different values of polytropic efficiencies, and by a diffuser with different values of total pressure drop across the diffuser expressed by pg /p,/ , respectively. The comparison reveals that for the same pressure gain pg/pl, the ideal shock compression efficiency may far exceed the efficiency obtained by a decelerating diffuser or a compressor. For example, it is seen that for a pressure gain of up to p2 /p, =2.2, ”Shock is greater than about 93% and thus is greater than those of typical diffusers and compressors. Therefore a gain in cycle performance can be expected when a compressor (or a diffuser) is replaced by an unsteady-flow device utilizing shock waves. Flow friction-effects would lower the efficiency of wave devices and reduces their efficiency advantage (not shown in Figure 2), but the relative advantage is expected to persist. Detailed calculations for developing Figure l and Figure 2 are presented in Appendix A. 1 :1 / / T. K $0510... 0'9 -. ._ ‘T ' " '- -= :v—Iu— “wt—v _. 0,85. / Diffuse/r , .. = " ‘ 7 e I / , ' ’ 5 08 I / / " I g ' T /' mf A_ L_..__~__~ Iii”... .. j g l I / ea — A i 7’ T a I ’ / 9. E 0.7 at. I, / 0 l I _ v, - ._ _ I I I » A - —- - I I I i ‘-:-.-'rr,-r.-a;‘ E7331“. €':‘:"",' 0.6 -' j l I I I III 0.5 III I l ' ' ‘ 1 1.5 2 2.5 3 3.5 4 mm: Figure 2: Shock wave, diffuser, and compressor isentropic efficiencies as functions of pressure gain 1.2 Wave Rotors There are at least two classes of wave machines: rotating wave machines and non- rotating wave machines. Dynamic pressure exchangers and wave rotors are two types of rotating wave devices developed to reach the high performance targets of thermodynamic cycles. In Europe, however, both terms have been often used interchangeably [3]. The essential feature of wave rotors is an array of channels arranged around the axis of a cylindrical drum. As schematically shown in Figure 3, the drum rotates between two end plates each of which has a few ports or manifolds, controlling the fluid flow through the channels. The number of ports and their positions vary for different applications. By carefully selecting their locations and widths to generate and utilize wave processes, a significant and efficient transfer of energy can be obtained between flows in the connected ducts. Stator end plate Channels Rotating drum Figure 3: Schematic configuration of a typical wave rotor Through rotation, the channel ends are periodically exposed to the ports located on the stationary end plates initiating compression and expansion waves within the wave rotor channels. Thus, pressure is exchanged dynamically between fluids by utilizing unsteady pressure waves. Therefore, unlike a steady-flow turbomachine which either compresses or expands the fluid, the wave rotor accomplishes both compression and expansion within a single component. To minimize leakage, the gap between the end plates and the rotor has to be very small or the end plates with sealing material could contact the rotor. An inverted design with stationary rotor and rotating ports is also possible [4]. Such a configuration has more than one rotating part and usually doubles interfaces that need to be sealed between rotating and stationary parts. However, it may be preferred for laboratory investigations because it easily enables flow measurement in the channels where the important dynamic interactions take place. However, this arrangement rarely seems to be convenient for commercial purposes. The rotor may be gear or belt driven or preferably direct driven by an electrical motor (not shown). The power required to keep the rotor at a correctly designed speed is negligible [5, 6]. It only needs to overcome rotor windage and friction in the bearings and contact sealing if used. Alternatively, rotors can be made self-driving. This configuration, known as the “free-running rotor”, can drive itself by using the momentum of the flow to rotate the rotor [7, 8]. In wave rotor machines, two basic fluid-exchange processes usually happen at least once per revolution of the rotor: the high-pressure process (charging process) and the low-pressure process (scavenging process). In the high-pressure process, compression waves transfer the energy directly from a fluid at a higher pressure (driver fluid) to another fluid at a lower pressure (driven fluid). In the low-pressure process the driver fluid is scavenged from the rotor channels, using expansion waves. Generation of expansion waves allows ingestion of a fresh low-pressure fluid into the rotor channels. There are several important advantages of wave rotor machines. Their rotational speed is low compared with turbomachines, which results in low material stresses. But they can respond on the timescale of pressure waves, with no rotor inertial lag. From a mechanical point of view, their geometries can be simpler than those of turbomachines. Therefore, they can be manufactured relatively inexpensively. Also, the rotor channels are less prone to erosion damage than the blades of turbomachines. This is mainly due to the lower velocity of the working fluid in the channels, which is about one-third of what is typical within turbomachines [5]. Another important advantage of wave rotors is their self-cooling capabilities. They are naturally cooled by the fresh cold fluid ingested by the rotor. Therefore, applied to a heat engine, the rotor channels pass through both cool air and hot gas flow in the cycle at least once per rotor revolution. As a result, the rotor material temperature is always maintained between the temperature of the cool air which is being compressed, and the hot gas which is being expanded. Despite very attractive features, several challenges have impeded the vast appearance of commercial wave rotors in some applications though numerous research efforts have been carried out during the past century. Besides unusual flow complexity and anticipated off-design problems and uncertainties about the selection of the best wave rotor configuration for a particular application, the obstacles have been mainly of a mechanical nature, like sealing and thermal expansion issues, as mentioned throughout this study. However, due to the recent energy crises, technology improvement, and economic reasons, new desires for wave rotor technology have been stimulated. 1.3 Wave Rotor Applications As a combined expansion and compression device, the wave rotor can be used as a supercharging device for IC engines, a topping component for gas turbines, in refrigeration cycles, and more. In advanced configurations, combustion occurs internally in the wave rotor channels allowing extremely short residence times at high temperature, hence potentially reducing NOx emissions. A condensing wave rotor may be viewed as a similarly advanced configuration that enhances the performance of water refrigeration cycles. Recently, wave rotor technology has been envisioned to enhance the performance of ultra-micro gas turbines manufactured using microfabrication technologies. In this study, the focus is on using wave rotor as topping components for gas turbine cycles. 1.3.1 Gas Turbine Applications Over the last decades, gas turbines have played a major role in a wide size range of propulsion and power generation systems. A schematic diagram for a simple-cycle, single-shafl gas turbine is shown in Figure 4. It consists of three components: compressor, combustion chamber, and turbine. Combustion Chamber Compressor Turbine Figure 4: Schematic configuration of a simple gas turbine engine Air enters the compressor at state “0” and is compressed to some higher pressure. Upon leaving the compressor, the compressed air enters the combustion chamber at state “1” where fuel is injected in a nearly constant pressure process. At state “4”, the hot gases enter the turbine where part of their thermal energy is converted into work. In a power generation application, some of the work produced by the turbine is used to drive the compressor, and the remainder is available as output work produced by the cycle. The expanded gases leave the turbine to the surroundings at state “5”. There are several methods to enhance the performance of gas turbines, including improvement in aerodynamic designs of turbomachinery components or thermodynamic cycle enhancement. The aerodynamics of turbomachinery has already yielded very high component efficiencies up to around 90% [9]. Further improvement is possible, but huge gains seem unlikely. From a thermodynamic point of view, increasing the turbine inlet temperature is the most efficient way to improve both thermal efficiency and output work. For a given engine, this can be achieved by increasing the cycle pressure ratio using a larger compressor. This is shown in a schematic temperature-entropy (T-s) diagram in Figure 5, where an increased pressure ratio has modified the baseline cycle from O-lb-4b-5b to 0-1-4-5. The numbers on this diagram correspond to the numbers used in Figure 4. According to the first law of thermodynamics, the cycle net work (Wm) is equal to the cycle net heat transfer. Since the cycle with the higher pressure ratio has a greater heat supply (QH) with the same heat rejected (QL) as the baseline cycle, it has greater output work and also greater thermal efficiency defined by: Wnet : QH —QL :1_& QH QH QH ’7: (1) However, the maximum temperature of the gas entering the turbine is usually fixed by material considerations. Therefore, cycle 0-1-4-5 has an intolerably high turbine inlet temperature (T4) much greater than that for the baseline engine (T 41,). One innovative solution for this turbine blade temperature limitation is to cool the burned gases by extracting energy from the flow before it enters the turbine. This way, the cycle peak temperature is de-coupled from the turbine inlet temperature. It can be accomplished by letting the hot gas leaving the combustion chamber compresses the air coming from the compressor, utilizing shock waves in an appropriate geometry. This is the basic principle behind the operation of wave rotors. 0-1-4-5 Higher pressure ratio cycle 0-1b-4b-5b Baseline cycle T turbine Temperature Entropy Figure 5: Schematic temperature-entropy diagram for a gas turbine with low and high pressure ratios 0 Wave Rotor Topping Cycles In a wave-rotor-topped cycle, the combustion can take place at a higher temperature while the turbine inlet temperature can be equal to that of the baseline cycle. Also, a pressure gain additional to that provided by the compressor is obtained by the wave rotor. Thus, the wave rotor can increase the overall pressure ratio and peak cycle temperature beyond the limits of ordinary turbomachinery. As a result, the performance enhancement is achieved by increasing both the thermal efficiency and the output work, hence reducing the specific fuel consumption rate considerably. In the following, a four—port wave rotor integrated into a simple gas turbine cycle is selected to prove these facts and its principal operation is briefly discussed In a conventional arrangement, the wave rotor is embedded between the compressor and turbine “parallel” to the combustion chamber. Figure 6 illustrates how a four-port wave rotor is used to top a gas turbine cycle. Figure 7 schematically shows how the wave rotor can be embedded physically into a baseline engine that uses single stage radial compressor and turbine. Combustion Chamber b.) 0 Wave Rotor 1‘ Compressor Turbine Figure 6: Schematic of a gas turbine topped by a four-port wave rotor \Vm'c Rotor Turhlne Combustor Compressor Figure 7: Schematic example of the physical implementation of a wave rotor in a gas turbine (exploded view, piping not shown) 10 Following the flow path shown in Figure 6, air from the compressor enters the wave rotor (state 1) and is further compressed inside the wave rotor channels. After the additional compression of the air in the wave rotor, it discharges into the combustion chamber (state 2). Here, combustion takes place at a higher pressure and temperature than in the baseline engine. The hot gas leaving the combustion chamber (state 3) enters the wave rotor and compresses the air received from the compressor (state 1). To provide the energy transfer to compress the air, the burned gas expands and is afterward scavenged toward the turbine (state 4). Due to the pre-expansion in the wave rotor, the burned gas enters the turbine with a lower temperature than that of the combustor exit. However, the gas pressure is still higher than the compressor exit pressure by the pressure gain obtained in the wave rotor. The turbine inlet total pressure is typically 15 to 20% higher than the air pressure delivered by the compressor [10]. This pressure gain is in contrast to the untopped engine, where the turbine inlet pressure is always lower than the compressor discharge pressure due to the pressure loss across the combustion chamber. As a result of the wave rotor pressure gain, more work can be extracted from the turbine increasing overall engine thermal efficiency and specific work. Finally, the channels are re- connected to the compressor outlet, allowing fresh pre-compressed air to flow into the wave rotor channels and the cycle repeats. The general advantage of using a wave rotor becomes apparent when comparing the thermodynamic cycles of baseline and wave—rotor—enhanced engines. Figure 8 shows schematic T-s diagrams of the baseline engine and the corresponding wave-rotor-topped engine. The shown wave rotor implementation is the one most commonly discussed in references, referred to as Case A in this study. It is evident that both gas turbines are 11 operating with the same turbine inlet temperature and compressor pressure ratio. Each wave rotor investigated in this work has zero shafi work. Therefore, the wave rotor compression work is equal to the wave rotor expansion work. Thus, the energy increase from state “lb” to “4b” in the baseline engine and from state “1 ” to “4” in the wave-rotor- topped engine is the same. This results in the same heat addition for both cycles. However, the output work of the topped engine is higher than that of the baseline engine “t” due to the pressure gain across the wave rotor (p,4>p,4b , where subscript indicates total values). Therefore, the thermal efficiency for the topped engine is higher than that of the baseline engine. The inherent gas dynamic design of the wave rotor compensates for the combustor pressure loss from state “2” to “3”, meaning that the compressed air leaving the wave rotor is at higher pressure than the hot gas entering the wave rotor. 0-1-2-3-4-5 Engine topped with a wave rotor Tturbine — —————————————— - — —-. E 3 ‘5 I- Q) Q. E O [— 1b= 0 0-1b-4b-5b Baseline engine Entropy Figure 8: Schematic T-s diagrams for a gas turbine with and without a wave rotor 12 Other advantageous implementation cases for the wave rotor into the given baseline engine are also possible. Four more advantageous cases have been extensively studied in this work and their advantages and disadvantages will be discussed in detail in Chapter 4. The goal of this chapter was to introduce the reader to the concept of the wave rotor and its application for gas turbine cycles. Next chapter (Chapter 2) provides a succinct review of past and current research in developing wave rotor technology. Chapter 3 describes the gasdynamic principle behind the operation of wave rotors. In Chapter 4, a comprehensive and systematic performance analysis of two microturbines known as the C-30 and C-60 engines which are topped with a four-port wave rotor in various wave- rotor—topping cycles is presented. Chapter 5 describes an analytical design procedure for the critical high-pressure phase of four-port wave rotors to predict some of their useful design parameters. A CF D simulation is also performed showing that commonly available CFD software FLUENT can be utilized for wave rotor simulations and to support their design. Finally, Chapter 6 introduces and studies several innovative wave rotor concepts and designs including radial-flow wave rotors, integrating wave rotors in ultra-micro gas turbines (UuGT), and water refrigeration systems working with wave 1'0t01’S . 13 CHAPTER 2: HISTORICAL REVIEW OF WAVE ROTOR TECHNOLOGY Recent advances and experiences obtained by the wave rotor community have renewed interest in this technology. These advances include new computational capabilities allowing accurate simulation of the flow field inside the wave rotor, and modern experimental measurements and diagnostic techniques. Improvements in aerodynamic design, sealing technologies, and thermal control methods have been sought. Recent developments in related unsteady flow and combustion processes of pulsed detonation engines have also provoked renewed interest. For this reason it is worthwhile to review the past and current work.1 The Early Work (1906-1940) The earliest pressure exchanger was suggested by Knauff in 1906 in which he did not employ the action of pressure waves [11]. The pressure exchanger introduced by him consisted of a cellular drum that rotates between two end plates containing several ports through which flows with different pressures enter and leave, exchanging their pressure. Knauff initially described rotor channels with curved blades and proposed inclined nozzles in the stator to achieve output shaft power (pressure exchange engine) besides pressure equalization inside the rotor. Reported by Pearson [12], Knauff in his second patent in 1906 [13] and Burghard in 1913 [14] proposed a simpler device in which the pressure exchange takes place in long narrow channel configurations @ressure exchanger) known later as the Lebre machine following Lebre’s patent in 1928 [15]. Today, the term “static pressure exchanger” is normally given to this type of device. Around 1928, Burghard proposed the utilization of pressure waves in another invention ' Materials presented in this chapter have been accepted for publication in 2004 International Mechanical Engineering Conference, ASA/{E Paper [MEC52004-60082, USA, Nov 2004. 14 [16] that was termed the “dynamic pressure exchanger” to distinguish it from the static pressure exchanger. Here, the term “dynamic” implies the utilization of pressure waves in both compression and expansion processes taking place inside the rotor channels. However, difficulties mainly related to poor knowledge about unsteady flow processes limited the dissemination of the dynamic pressure exchanger concept [3] until World War II when Seippel in Switzerland implemented this concept into real engines, as discussed below. The Comprex® Pressure Wave Supercharger (l940-Present) Brown Boveri Company (BBC), later Asea Brown Boveri (ABB) and now Alston, in Switzerland has a long history in wave rotor technology. As reported by Meyer [17], their initial investigations in the beginning of the 19405 were aimed at implementing a wave rotor as a topping stage for a 1640 kW (2200 hp) locomotive gas turbine plant of British Railways [18-21]. They expected to obtain a power increase of 80% (2983 kW or 4000 hp) and a 25% efficiency increase (from 18% to 22.5%) based on the patents of Seippel [22-25]. This arrangement is shown in Figure 9. The wave rotor had 30 channels rotating at 6000 rpm, with two opening ports on each side through which air and gas entered and left. It had originally shown a pressure ratio up to 3:1 and total efficiency of 69% in previous tests during 1941-1943, which could approximately result in a 83% efficiency for each compression and expansion process [17]. 15 " Wave Rotor fl! Figure 9: Wave rotor as a topping stage for the locomotive gas turbine, taken from Ref. [17] The first wave rotor worked satisfactorily, proving the concept of wave rotor machines. However, its performance when installed in the engine was far from expectations, mainly because of its inefficient design and crude integration [20]. Seippel’s work also initiated the notion of using the wave rotor as a pressure wave supercharger for diesel engines. The extensive practical knowledge accumulated by BBC during investigations of gas turbine topping cycles was then used to develop pressure wave superchargers first by the ITE Circuit Breaker Company in the US [26-28]. In an effort jointly sponsored by the US Bureau of Aeronautics and ITE supervised by Kantrowitz of Cornell University and Berchtold of ITE, the first units were successfully manufactured and tested on vehicle diesel engines between 1947 and 1955. As a result of this success, a co-operative program with BBC was started in 1955 and BBC decided to concentrate on the development of pressure wave superchargers for diesel engines, due to their higher payoff compared to other applications [29]. As a manufacturer of superchargers, BBC later continued the project in collaboration with the Swiss Federal Institute of Technology (ETH Zurich). While the first prototype was installed in a truck engine in 1971 [30], the supercharging of passenger car diesel engines was started in 1978 [31, 32] with a first successful test on an Opel 2.1 liter diesel engine [32, 33]. This supercharger was given the trade name Comprex® shown in Figure 10. The port arrangement indicates the use of two operating cycles per revolution, shortening the rotor length and reducing thermal loads. The main advantage of the Comprex® compared to a conventional turbocharger is its rapid response to changes in engine operating conditions where the turbocharger is less responsive, due to its rotational inertia. Furthermore, as the efficiency of the Comprex® is independent of scale, its light weight and compact size make this device attractive for supercharging small engines (below about 75 kW or 100 hp) [34, 35]. Figure 10: The Comprex®, taken from Ref. [34] By 1987, the first wide application of the Comprex® in passenger cars occurred in the Mazda 626 Capella [7, 36]. Since then, ABB’s Comprex® pressure wave supercharger has been commercialized for several passenger car and heavy diesel engines. For instance, once Mazda produced 150,000 diesel passenger cars equipped with pressure wave superchargers [37]. The Comprex® has also tested successfully on vehicles such as the Mercedes-Benz diesel car [8], Peugeot [29], and Ferrari [29]. The progress by BBC/ABB took almost five decades to accomplish. The successful development of the Comprex® has been enabled by efforts of numerous researchers. Besides the above mentioned names, only some more are listed here: Gyarrnathy [6], Burri [38], Wunsch [39], Croes [40], Summerauer [41], Kollbrunner [42], Jenny [43], Keller [44], Rebling [45], and Schneider [46]. Further references related to the development of the Comprex® by BBC [47-60] and other organizations [61-71] until 1990 can be found in the literature. By the end of the 19805, when the Comprex® activity was transferred to the Mazda company in Japan [3, 72], researchers at ABB turned to the idea of utilizing wave rotor technology for gas turbine applications [9, 73]. During 19908, a few groups continued the development of pressure wave superchargers. Nour Eldin and his associates at the University of Wuppertal in Germany have developed a fast and accurate numerical method for predicting the unsteady-flow field in pressure wave machines, using the theory of characteristics [74-80]. Piechna et al. at the Warsaw University of Technology in Poland have developed experimentally validated one-dimensional and two-dimensional numerical codes to analyze the flow field inside the Comprex® [81-87]. Piechna has also proposed a compilation of the pressure exchanger with the internal combustion wave rotor, presenting the idea of the 18 autonomous pressure wave compressor [88]. Oguri et al. at Sophia University in Japan have performed measurements on a car gasoline engine supercharged by the pressure wave supercharger to investigate the feasibility of such integrations [89]. This effort sought to extend the application from diesel engines to gasoline engines and achieved a satisfactory increase of thermal efficiency of the supercharged engines. Guzzella et a1. [35, 90-94] at ETH in Switzerland have developed a control-oriented model that describes the entire engine supercharged by pressure wave devices, with special emphasis on the modeling of transient exhaust gas recirculation phenomena. The experimentally validated model has introduced an optimized strategy to operate a supercharged engine with good drivability. Finally, an investigation of Comprex® supercharging on diesel emissions has been recently performed in Turkey [95], demonstrating that the Comprex® has the potential for reducing NOx in diesel engines. To date, the Comprex® has been recognized the most successful application of wave rotor technology and represents a practical utilization of the wave rotor concept. The Comprex® development by BBC/ABB also has established fabrication techniques for wave rotors in commercial quantities and considered as a matured and reliable machine for internal combustion engine supercharging. For this application, BBC/ABB has solved difficult development challenges like sealing against leakages, noise, and the thermal stress problems. For instance, enclosing the rotor in a pressurized casing and using a rotor material with a low thermal expansion coefficient over the operating temperature range has kept Comprex® leakages to an acceptable level [29]. Furthermore, several pockets have been cast into the end plates to control wave reflections and to achieve good off- design performance when engine speed changes [59]. 19 In recent years, Swissauto WENKO AG in Switzerland has developed a modern version of the pressure wave supercharger [37]. This new generation of Comprex® known as the Hyprex® is designed for small gasoline engines. It benefits from new control features, enabling higher pressure ratios at low engine speeds, further reduced noise levels, and improvement of the compression efficiency at medium or high engine speeds. The Hyprex® has been successfully demonstrated in the SmILE (Small, Intelligent, Light and Efficient) vehicle which is a modified Renault Twingo, achieving very low specific fuel consumption and low emissions. ETH is collaborating closely in this effort by developing control systems for the proper operation of the device. Cornell Aeronautical Laboratory and Cornell University (1948-2001) Inspired by the cooperation with BBC in the late 19405, work on unsteady-flow concepts was initiated at Cornell Aeronautical Laboratory (CAL). Among several novel concepts including development of energy exchangers for gas turbine cycles and various stationary power applications [96], the CAL Wave Superheater was built in 1958 and utilized until 1969 [21]. The 2 m diameter wave superheater used heated helium as the low molecular weight driver gas to provide a steady stream of high-temperature and high- pressure air for a hypersonic wind tunnel test facility. It compressed and heated air to more than 4000 K and up to 120 atm for run times as long as 15 seconds. Figure 11 is a photograph of this device. The CAL Wave Superheater was a landmark demonstration of the high temperature capabilities of wave rotor devices [21, 96]. 20 Figure 11: Photograph of the CAL Wave Supercharger, taken from Ref. [5] Around 1985, Resler, a former member of the CAL Wave Superheater team, resumed the wave rotor research at Cornell University. His efforts and those of his group led to the development of new wave rotor concepts and analytic methods for three-port wave rotor diffusers [97], double wave rotor cycles [98], five-port wave rotors [98—105], and supersonic combustor aircraft engines using wave rotors [106]. Five-port wave rotors have shown significant potential for reducing NOx in gas turbine engine applications. Figure 12 illustrates a double wave rotor in a gas turbine cycle. The idea of using a compound unit consisting of two (or multiple) wave rotors, one supercharging the other, is also reported in an early German patent by Muller in 1954 [107], as stated by Azoury [34]. 21 X Bypass / ' 2 / High Low \ Compressor Pressure N Pressure # Turbine Turbine ' Figure 12: Schematic of a double wave rotor cycle, taken from Ref. [98] Power Jets Ltd (1949-1967) In parallel with but independent of Seippel’s efforts in 19403, Jendrassik, former chief engineer of the Gantz Diesel Engine Company of Budapest, was working on the development of wave rotor machines for gas turbine applications [20, 108—110]. He developed one of the first concepts for wave rotor applications to aircraft engines, proposing the wave rotor as a high pressure topping stage for early aircraft engines [111, 112]. His ideas stimulated the govemment-controlled company of Power Jets Ltd in the UK to become active in the wave rotor field in 1949. Even though the initial intent of Power Jets Ltd was to use wave rotor technology for IC engine supercharging, the interest was later extended to several other applications including air cycle refrigerators, gas turbines, pressure equalizers, and dividers [3, 5, 20, 110]. For instance, two prototype air-cycle refrigerators using wave rotors were designed and employed in gold mines in India and South Africa for environmental cooling purposes. They performed the same duty as equivalent vapor-cycle machines, but with lower weight and bulk. After Jendrassik's death in 1954, theoretical and experimental work continued at Imperial 22 College, University of London, directed by Spalding and Barnes and also by Ricardo Company in the UK [20, 113]. The experimental divider test rig at Imperial College is shown in Figure 13. Detailed information related to Power Jets Ltd efforts can be found in company reports listed in Ref. [5]. Figure 13: The experimental divider test rig at Imperial College, taken from Ref. [5] Most of these efforts were experimental and thus expensive. Computational methods and digital computer facilities were too poorly developed in that time, so extensive theoretical methods required to improve the progress were too difficult. Cycle analyses by hand calculations were tedious and impractical. Spearheading the development of CFD methods, Spalding of Imperial College formulated a numerical procedure for wave rotors considering the effects of heat transfer and friction. It utilized novel features to ensure solutions free from instabilities and physical improbabilities [20]. Based on this numerical model, a computer program was developed by Jonsson [114] and it was 23 successfully applied to pressure exchangers [1 15—117]. Spalding’s students, Azoury [1 18] and Kentfield [119], continued their efforts on different theoretical aspects of pressure exchangers [3, 5, 20, 34, 110, 117, 120-123] despite the dissolution of Power Jets Ltd in 1967 [20]. Ruston-Hornsby Turbine Company: The Pearson Rotor (Mid 19505 - 1960) In the U.K. of the mid 19505, besides the work at Power Jets Ltd and Imperial College, the Ruston—Homsby Turbine Company, manufacturer of diesel engines and industrial gas turbines, supported the construction and testing of a different kind of wave rotor designed by Pearson [124, 125]. This unique wave rotor, known as the wave turbine engine or simply the wave engine, has helical channels that change the direction of the gas flows producing shaft work similar to a conventional turbine blade. With the financial support by the company, Pearson designed and tested his wave rotor in less than a year, shown in Figure 14. Figure 14: The Pearson rotor (left) and rear and front stator plates (right), taken from Ref. [124] The rotor has a 23 cm (9") diameter and a 7.6 cm (3") length. The engine worked successfully for several hundred hours in a wide range of operating conditions (c.g., 3000-18000 rpm) without variable porting, and produced up to 26 kW (35 hp) at its 24 design point with a cycle peak temperature of 1070 K and a thermal efficiency of around 10%. While the performance results were slightly less than the expected design performance (mainly due to the combined effect of excessive leakage and incomplete scavenging), higher performance seemed to be possible with more careful design and development. The design of the engine was based on many wave diagrams using the method of characteristics that accounted for all internal wave reflections. The engine utilized extra ports and injection nozzles to control and cancel unwanted reflected waves. The engine had a length of only one third of its diameter despite having only one cycle per revolution [12]. The sealing and bearings were carefully adapted considering rotor thermal expansion. Unfortunately, the engine was accidentally wrecked due to over speeding from an improperly connected fuel line while the company was suffering financial difficulties. Despite the success achieved, the wave engine experiments were far from the norm of ordinary projects at Ruston-Homsby and it was considered a redundant project. Tragically, the wave rotor project was canceled when success seemed so close. Further efforts by Pearson to attract additional funding by other sources to commercialize his engine were unsuccessful. In the history of wave rotor technology, the Pearson rotor and the Comprex® are known as the most successful wave rotor machines developed to date [20, 29, 109]. Both devices have worked efficiently over a wide range of operating conditions, demonstrating good off-design performance although the less-publicized CAL Wave Superheater was an equal success. Nevertheless, the Pearson rotor is a notable wave rotor for producing a significant power output in addition to being a successful pressure exchanger. 25 General Electric Company (1956-1963) While Pearson was developing his novel wave engine, General Electric Company (GE) in the US initiated a wave rotor program in 1956 [126]. The work was motivated by earlier work at NASA Langley initiated by Kantrowitz and continued by Huber [127] during the development of a wave engine in the early 19508 and later in 1954-1956 developing pressure gain combustors (constant volume combustion) [126]. GE studied a new configuration of wave rotor in which combustion took place inside the rotor channels (internal combustion wave rotors). Such an arrangement eliminates the need for an external combustion chamber used in the gas turbine cycle, resulting in a significantly lower weight, less ducting, and a compact size. In the period of 1956 to 1959, the methods used at NASA were analyzed, improved and applied to the design and fabrication of the first internal combustion wave rotor demonstrator. As reported by Weber [2], the test rig was first tested at the California Advanced Propulsion Systems Operation (CAPSO) of GE. After 20 seconds of operation, due to the heating and expansion phenomena, the rotor seized between the end plates causing an abrupt stop. The test demonstrated the difficulty of clearance control between the end plates and rotor during thermal expansion. Rotor expansion is an especially challenging problem in the design of wave rotors. While the running clearance between the end plates and rotor must be kept as small as possible, the rotor tends to expand thermally due to hot gases in the rotor. Henceforth, GE resorted to inferior rubbing type seals, and tested only pressure- exchange configurations from 1960 to 1961 [126]. Despite flow leakage, due to the poor sealing method, respectable wave rotor overall pressure ratios of 1.2 to 1.3 were achieved. Meanwhile, a feasibility study was initiated for reducing compressor stages of a T-58 GE-06 engine by using a wave rotor. It showed a considerable reduction in overall 26 engine weight and cost, and a 15% reduction in specific fuel consumption rate. These results motivated GE further to come up with a conceptual design layout of such an advanced engine. However, further rig experiments revealed other mechanical and aerodynamic shortcomings including start-up, bearing durability, fuel system complications and control [10]. GE also pursued designing a shaft work output wave rotor. Over the period from 1961 to 1963, Klapproth and his associates at GE in Ohio fabricated and tested a wave engine using air-gap seals. An ideal wave diagram of this engine is shown in Figure 15. The engine worked continuously, but it did not produce the anticipated net output power. It is believed that insufficient attention was given to account for internal wave reflections, thus, the flow field calculations were inaccurate [29]. Simplifications were unavoidable at that time and generation of wave diagrams by hand required considerable time and effort and small design changes necessitated a lengthy recalculation. Although the Klapproth rotor did not produce the expected performance, it clearly demonstrated the possibility of the complete exchange of energy within the wave rotor. GE development of the wave rotor was canceled in 1963 due to shifting funds from turbine engine development to space exploration and rocket propulsion [2], and GE’s commitment to pursue large engine development exclusively [126, 127]. 27 BUFFER . -F f 4475!? — 5 330— / \2 300—. Q Q - Q 270—. S i N * Q 1111 74 110—. Ilncr for;- Happy: LAYER l \\\\\\\\\\\\\\ /////////777 I//"’7//////7 Figure 15: Ideal wave diagram of the Klapproth rotor, taken from Ref. [126] 28 PORT DELI VERY our/14265 Paar General Power Corporation (Mid 1960s - 1984) In the mid 19605, General Power Corporation (GPC) started a wave rotor program originally intended for a road vehicle engine application [29]. Over a period of about 20 years, GPC spent considerable time and money to successfully design and develop wave rotors. The work was initially supported by Ford Motor Company and later by the Department of Energy (DOE) and the US Defense Advanced Research Program Projects Agency (DARPA). Unfortunately, the GPC work is poorly documented making it difficult to further report about their wave rotor. Some information is briefly reported in Ref. [128]. As stated by Taussig [29, 109], while the GPC rotor shared some of the features of the Klapproth and Pearson rotors, it differed in several aspects taking into account its own unique design and operation. Figure 16 illustrates an ideal wave diagram of the GPC rotor, intended to produce reactive shafi power utilizing curved blades. 29 30: Cum Eoc 50:3 :98 0&0 2: Co Seaman 03:5 :83 n3 oeswi BO 8an :eumancO “:0 82.5 seam—5:50 ~593an 208—an0 @995me Etna \Illikfiij 30 .._< commoEEeU \ r 5 c2 58... II'III I- 5 893g Exam—5:80 cocanaxm its; 5 8an Scum—5:80 30 There have been reported several difficulties with the GPC design which resulted in poor output power. Its performance suffered from excessive blade curvatures, lack of control of reflected waves within the device required to make the wave system periodic within one revolution, and the absence of any strong impulsive loading of the rotor from inlet manifolds to produce shaft work. The latter was in contrast with the Pearson rotor that relied heavily on impulsive loading of the rotor blades to achieve power output. Furthermore, the GPC rotor had inadequate control on maintaining high performance for off-design operating conditions. Although GPC developed a computer code to obviate manual wave pattern design, accurate calculations were still tedious. Unfortunately Ford ultimately withdrew its support from the wave rotor research before completion of tests, mainly due to commitments to other engine development programs [129]. As a result, GPC discontinued development of the wave engine in the early 19805. Rolls-Royce (1965-1972) In the mid 19605, Rolls-Royce (R) in the UK began numerical and experimental wave rotor research [29]. BBC cooperated with R in the development of pressure exchange wave rotors as topping spools in gas turbine applications [131], with Berchtold of the ETH and Spalding of Imperial College serving as consultants [3]. Considerable efforts were made to design a wave rotor as a topping stage for a small helicopter engine (Allison Model 250) [130]. The BBC-RR engine utilized a reverse-flow wave rotor incorporated into a single turbine cycle. This was somewhat different from the cycle suggested by Berchtold and Lutz [63] in BBC gas-turbine-topping investigations, which employed a through-flow wave rotor integrated with both low-pressure and high-pressure turbines. BBC’s interests in wave rotors at that time were mostly related to development of small gas turbines for passenger cars, beset by poor efficiencies at sizes of 100 kW and 31 smaller [21]. Similar to previous wave rotor efforts, rotor designs protracted manual design methods. While the enhanced engine operated nearly as predicated, measured data revealed low performance mainly due to leakage [29]. Other difficulties related to the start-up and control are reported [10]. The program was abruptly canceled in 1972 amidst severe company financial difficulties [131]. As stated by Kentfield [3], contemporaneous rapid progress in turbomachinery technology may have disfavored high-risk projects, both at RR and GE. Mathematical Science Northwest Inc. (1978-1985) In the late 19705 Mathematical Science Northwest Inc. (MSNW, later Spectra Technology Inc., and now STI Optronics Inc.) investigated various applications of wave rotors [21]. Under the sponsorship of DOE and DARPA, they considered a broad range of stationary power systems such as magnetohydrodynamic cycles (MHD) [29], combined cycles integrated with gasification plants [132], pressurized fluidized bed (PFB) power systems [133], and also propulsion and transportation applications [109]. Furthermore, significant numerical and experimental efforts were performed developing a laboratory wave rotor known as the MSNW wave rotor [134-137], shown in Figure 17. With diameter of 45 cm, it consists of 100 channels each with a 40 cm length. It is a four- port wave rotor with two additional small ports provided to cancel pressure waves at critical rotor locations providing more uniform port flows and a higher transfer efficiency [137]. The design rotor speed is reported as 1960 rpm and a pressure ratio of approximately 2.5 was achieved. A measured wave rotor efficiency of 74% is reported [138]. Besides successful tests using several configurations (clearance variations, port sizes, etc.) and various operating conditions, experiments were designed to verify the scaling laws for predicting the performance of larger machines [132]. 32 Figure 17: The MSNW experimental set up, taken from Ref. [21] The MSNW wave rotor was initially designed based on the method of characteristics using only the Euler equations, but later a one-dimensional unsteady computer code (the FLOW code) was used for optimizations and modifications of the MSNW design [109]. The modifications led to improvement in obtaining a very good agreement between the numerical and experimental results in a wide range of operating conditions. Analytic estimations involving the ideal wave patterns also supported the obtained results. The FLOW code which was developed specifically for both pure pressure exchanger wave rotor and wave engine analyses, uses the flux-corrected transport algorithm solving Euler equations accounting for heat transfer, viscosity, gradual port opening, and flow leakage. The sensitivity of wave rotor performance to tip speed, port placement and size, inlet and outlet flow conditions, channel geometry, number of channels, leakage, and heat transfer was analyzed for both on-design and off-design conditions. For instance, it was 33 outlet flow conditions, channel geometry, number of channels, leakage, and heat transfer was analyzed for both on-design and off-design conditions. For instance, it was concluded that heat transfer losses were negligible and leakage was recognized as a key problem for efficient wave rotor operation. Numerical work has been also reported for a nine-port wave rotor concept to resolve the problem of nonuniform port flows and poor scavenging. MSNW has used the knowledge obtained through their investigation to establish preliminary wave rotor designs for a small turbofan engine generating 600 lb thrust at sea level condition [109, 138]. A conceptual design of such engine integrated with a pressure exchanger wave rotor is illustrated in Figure 18. Performance calculations for both on- design and off-design flight conditions using a cycle performance code and the FLOW code simulation have predicted significant performance improvements of such an enhanced engine. No new material development for such combined engines was required, proving the possibility of designing such engines by using available technology. Despite all successful achievements and satisfying results, the wave rotor activity at MSNW was discontinued in the mid 19803. No specific reasons for this cancellation are reported. 34 5* « .- 55,334.; ' Conceptual design of a turbofan engine incorporating a wave rotor, taken from Ref. [138] Figure 18 Naval Postgraduate School (1981-1986) In 1981, the Office of Naval Research (ONR) agreed to monitor a joint DARPA/ONR program to evaluate the wave rotor concept and its potential application in propulsion systems [127]. Following this decision, Turbopropulsion Laboratory (TPL) at Naval Postgraduate School (NPS), directed by Shreeve, started an extensive numerical and analytical wave rotor program. To support the accuracy of the computational results, the wave rotor apparatus formerly used by Klapproth at GE was transferred to TPL and some preliminary tests were carried out. It is reported that the rotor produced some shaft work running at approximately 5000 to 6000 rpm [139]. No further experimental details are reported. For numerical simulations, two different approaches to the solution of the unsteady Euler equations were examined in the overall program. First, Eidelman developed a two- dimensional code based on the Godunov Method to analyze the flow in wave rotor channels [140-143]. Unlike contemporary one-dimensional approaches [144], the two- dimensional code showed the effect of gradual opening of the channels. The main conclusion of these studies is that if the channels are straight, the flow remains nearly one-dimensional, which in turn leads to minimal mixing losses caused by rotational flow in the channels [145]. However, when the channel of the wave rotor is curved, even an instantaneous opening of the channel does not lead to the development of a one- dimensional flow pattern with small losses. Computation time of such a two-dimensional code has been reported to be quite long. For faster computations, a one-dimensional, first order time-accurate code was introduced by Mathur based on the Random Choice Method for solving the Euler equations [146, 147]. The unconditionally stable code, called WRCOMP (wave rotor component), calculated the unsteady-flow process inside 36 the wave rotor, inlet and outlet opening times and other useful design parameters required for a preliminary design. The outputs from WRCOMP are used in a second program, called ENGINE, for turbofan jet engine performance calculations [148-150]. The results confirmed the significant performance improvement expected by integrating a wave rotor into a turbofan engine. Work was planned toward incorporating the effects of friction and heat transfer into WRCOMP and also including other engine configurations in the ENGINE code. Some improvements to the WRCOMP code were later started [151, 152], but further development was not pursued after terminating the wave rotor research around 1986. It is also worth to point out that NPS sponsored the most comprehensive wave rotor conference in 1985 where worldwide participation in this conference took place [153]. The conference reviewed much of the history to that point. NASA Glenn Research Center (1990—Present) Since the late 19805, a sustained research program at NASA Lewis (now Glenn) Research Center (GRC), collaborating with the US Army Research Laboratory (ARL), Rolls-Royce Allison has aimed to develop and demonstrate the benefits of wave rotor technology for future aircraft propulsion systems [10]. In 1993, using a thermodynamic approach to calculate the thermal efficiency and specific power, Wilson and Paxson [154] published a feasibility study for topping jet engines with wave rotors. Applied to the case of an aircraft flying at Mach 0.8 , they have shown that a wave-rotor-topped engine may gain l...2% in efficiency and 10...16% in specific power compared to a simple jet engine with the same overall pressure ratio and turbine inlet temperature. Additionally, Paxson developed a quasi-one-dimensional gasdynamic model to calculate design geometry and off-design wave rotor performance 37 [155, 156]. The code uses an explicit, second order, Lax-Wendroff type TVD scheme based on the method of Roe to solve the unsteady flow field in an axial passage for time- varying inlet and outlet port conditions. It employs simplified models to account for losses due to gradual passage opening and closing, viscous and heat transfer effects, leakage, flow incidence mismatch, and non-uniform port flow field mixing. In order to verify wave rotor flow predictions and to assess the effects of various loss mechanisms [157, 158], a three-port wave-divider machine was constructed and tested [159—161] in a new wave rotor laboratory facility at GRC. Concurrently, the non-ideal behavior and losses due to multi-dimensional effects were studied by Welch [162-164] and Larosiliere [165-167]. Welch has also established macroscopic and passage-averaged models to estimate the performance enhancements of wave rotors [168, 169]. Based on experimental data, Paxson further improved the one-dimensional model [157, 158, 170, 171] and used it to evaluate dynamic behavior, startup transients, and channel area variation [172—175]. This model was then used as a preliminary design tool to evaluate and optimize a four-port wave rotor cycle for gas turbine topping [176]. This through- flow cycle was chosen based on several perceived merits, including relatively uniform rotor temperature, and the feasibility of integration with gas turbomachinery. As a result of these studies, a new four-port wave rotor was designed and built [177] to test the performance of this concept under scaled laboratory conditions. A photograph of NASA four-port wave rotor is shown in Fig. 16. However, a study by Rolls-Royce Allison discussed below indicated that thermal loads on the rotor and ducting predicted for the NASA wave rotor cycle in real engine conditions may be difficult to manage. In 38 response, Nalim and Paxson [178, 179] proposed an alterative cycle with a combustor bypass significantly lowering thermal loads. Figure 19: F our-port wave rotor of NASA Additional studies of the performance benefits of wave rotor topped gas turbines have been reported. In 1995, Welch et al. [180] predicted a 19.21% increase in specific power and a 16...17% decrease in specific fuel consumption compared with the baseline engines in performance calculations for small (300 to 500 kW) and intermediate (2000 to 3000 kW) wave-rotor-enhanced turboshaft engines. The same calculations for a wave- rotor-enhanced large turbofan engine, equal in thrust to the baseline engine, have shown a 6...7% reduction in thrust specific fuel consumption. Welch has also studied the possibility of curving the channels to create a wave turbine [181, 182]. In 1995, Nalim at NASA published a feasibility assessment of combustion in the channels of a wave rotor, for use as a pressure-gain combustor [183]. Combustion prediction capability was added to the wave rotor code by Nalim and Paxson [184], 39 enabling the exploration of wave cycles involving both detonation and deflagration modes of combustion. For uniform mixtures, a single reaction progress variable is utilized. Multiple species are represented for a variable fuel-air ratio in deflagration modes. Mixing controlled reaction is combined with a simple eddy diffusivity model. Other notable features that were incorporated are temperature kinetics factors and a simple total-energy based flammability limit [185]. The performance of detonative and deflagrative cycles was studied by combined CFD and system simulation. It was determined that deflagrative combustion with longitudinal fuel stratification could be accomplished over a reasonable time in wave rotors. The current NASA wave rotor research has been mostly focused on experimental tests with special attention to sealing technology [186-188], identified as a critical challenge in high-pressure wave rotor design. Rolls-Royce Allison (l990-Present) In 1996, Snyder and Fish [189, 190] of Allison Engine Company evaluated the Allison 250 turboshafi engine as a potential platform for a wave rotor demonstration, predicting an 18...20% increase in specific power and a 15.22% decrease in specific fuel consumption. They used a detailed map of the wave rotor cycle performance accomplished by Wilson and Paxson [10, 154, 176]. Allison (by now Rolls-Royce Allison) has also studied transition duct designs for integration with turbomachiney [191, 192]. This was later followed by investigations of pulse detonation wave rotors in the newly formed Allison Advanced Development Company (AADC). A novel four-port device is proposed [193] for supersonic turbofan engines [194], and was investigated in collaboration with Indiana University Purdue University Indianapolis (lUPUI) as discussed later. 40 University of Florida (1992-1998) Motivated by NASA wave rotor successes, Lear at the University of Florida initiated analytical and numerical methods to investigate different configurations of wave rotors. His team developed an unsteady two-dimensional numerical code using a direct boundary value method for the Euler equations to analyze the flow in wave rotors and their adjoining ducts, treating the straight or curved channel walls as constraints imposed via a body force term [195]. The code was later used to simulate the flow field of the three-port NASA wave rotor. They also introduced a preliminary design method for selecting the wave engine inflow and outflow blade angles. Furthermore, an analytical thermodynamic description of wave rotors was developed [196], which predicted potential increase in specific power of 69% and a 6.8% increase in thermal efficiency over a conventional gas turbine topped by a wave engine. A parametric study of gradual opening effects on wave rotor compression processes is reported, too [197]. ONERA in France (1995-1999) Fatsis and Ribaud at the French National Aerospace Research Establishment (ONERA) have investigated wave rotor enhancement of gas turbines in auxiliary power units, turboshaft, turbojet, turbofan engines [198, 199], accounting for compression and expansion efficiency, as well as mixing and pressure losses in the ducting. Their results show the largest gains and efficiency for engines with a low compressor pressure ratio and high turbine inlet temperature, such as turboshafi engines and auxiliary power units. These results are consistent with those obtained by NASA GRC [200]. They hve also developed a one-dimensional numerical code based on an approximate Rieman solver taking into account viscous, thermal, and leakage losses [198, 201], and applied it to three-port, through-flow, and reverse-flow configurations. 41 0 Recent Academic Work Besides ongoing research mainly at NASA, AADC, and ETH Zurich, a few universities have been conducting wave rotor research. To the knowledge of the author, the universities listed below are are active in this field. Purdue School of Engineering and Technology (1997-Present) Recent research at Indiana University Purdue University Indianapolis (IUPUI) by Nalim and coworkers has focused on internal combustion wave rotors, following the initial work at NASA described above. Deflagrative combustion with longitudinal fiJel stratification has yielded a wave rotor geometry competitive with pressure-exchanger designs using a separate combustor [185, 202]. Nalim has highlighted the importance of thermal management of leakage flows of rotor and end—wall temperatures, with illustration of the impact of the hot ignition gas and the cold buffer zones on the end walls. This is consistent with the major challenges revealed by the ABB experiment [73]. Radial stratification [203] using a pre-combustion partition has been proposed to introduce a relatively cooler buffer zone close to the leakage gaps, reducing hot gas or fuel leakage to the rotor cavity. Figure 20 is a contour plot of the temperature contour from a simulation of deflagrative combustion in a stoichiometric partition region propagating into a leaner mixture in the main chamber. Above and below the partitions, there is no fuel, and gas may leak out or in without danger of overheating or pre- ingnition. These thermal management approaches are possible utilizing extensive cycle design studies and analysis, and seek to alleviate the challenges previously recognized by ABB and NASA. This technique also helps burn leaner mixtures, resulting in reduced NOx emissions, similar to other pilot combustion or lean-bum techniques in conventional engines [204]. For this approach, radial leakage flows [205] and different combustion 42 models [206] have been studied in detail. These ideas have not yet been tested experimentally. Figure 20: Temperature distribution of partition exit flow, taken from Ref. [204] Detonative combustion cycles for propulsion engines have been also studied [207, 208]. Interest in detonative combustion initially focused on pulsed detonation engines (PDE) has evolved to the consideration of the wave rotor as an effective implementation of the concept [209], and a means of overcoming challenges to PDE concepts that involved integration with conventional turbomachinery. In effect the wave rotor provides automatic high-speed valving, nearly steady inflow and outflow, and the use of one or few steady ignition devices for multiple tubes. However, detonative combustion is fundamentally restricted to highly energetic mixtures and sufficiently large passage widths, and generates strong pressure waves. This results in the outflow being highly non-uniform in pressure, velocity, and possibly temperature. To better utilize the output of a wave rotor PDE, it has been proposed to add an ejector element to the wave rotor [210]. The rotary wave ejector admits bypass air after the detonation tubes to transfer energy and momentum. Numerical simulations using a quasi-one-dimensional code, modified to account for radial-type bypass flows, have shown that the specific impulse at static thrust conditions can be doubled, after accounting for flow-turning and shock losses, comparing with an equivalently loss-free PDE cycle. A sample wave diagram and 43 a schematic sketch are given in Figure 21, where the cold ejector gas flow is clearly distinguishable. Temperature Pressure Fuel Fraction 2.5 0 0.5 1 D 0 5 ‘l Posmon/Length Posrtion/Length Posttion/Lenglh Figure 21: Rotary Wave Ejector Pulse Detonation Engine, taken from Ref. [210] IUPUI has also investigated [211] the four-port detonation wave rotor proposed by AADC [193], in which a recirculation duct allows air that is compressed by the shock of a detonation wave to be reinjected with fuel. Air-buffer regions both between the fuel/air- combusted gas interface and at the exit end plate are inherent in the cycle design, allowing self-cooling of the walls. The inflow and outflow of this engine concept is designed to be nearly uniform and acceptable to modern turbines, compared to conventional rotary detonation cycles, as shown in Figure 22. 44 Transition duct Rotor Inlet endplate *0 tlll'llllt'le OI’ nozzle Transition duct from compressor or inlet ‘ Exit and plate with detonation initiation devices .. High pressure bufier air duets with fuel nozzles Figure 22: Wave Rotor PDE the ‘CVC’ Engine, taken from Ref. [211] A computational and experimental program is currently being conducted at IUPUI in collaboration with AADC to investigate the combustion process and performance of a wave rotor with detonative and near-detonative internal combustion [212]. A preliminary design method based on a sequence of computational models has been developed to design wave processes for testing in an experimental test rig. University of Tokyo (2000-Present) Nagashima et al. have developed one-dimensional [213] and two-dimensional [214] CFD codes to simulate the flow fields inside through-flow four port wave rotors, including the effects of passage-to-passage leakage. The codes have been validated with experimental data obtained by a single-channel wave rotor experiment Single-cell experiments can efficiently demonstrate the operation of actual wave rotor engines. The test rig, shown in Figure 23, consists of a stationary single tube, and two rotating plates connected to a shaft driven by an electric motor. This group has also explored the idea of 45 using wave rotors for ultra-micro gas turbines manufactured using microfabrication technology [215]. Driving Belt \ H Wfiullhlll‘l‘w i M I j ' is ‘ \ ‘ I .II I . Y 3) [ V r l l Oil Sea 1W“.”Iiimmiiw‘l u" vi «till . ll II I [I will I“ . , ., trill 1“ Figure 23: University of Tokyo single-channel test rig, taken from Ref. [214] In conclusion, Figure 24 summarizes the history of the wave rotor research reviewed here. The goal of this review was to report the continued interest in wave rotor technology and its wide variety of applications. Some of the latest efforts were discussed in more detail, inspiring further research and development on this topic 46 ‘———EUROPE —'—-—> U.S. Japan Germany France Poland U.K. Switzerland General Electric Comel Univ. ll ITE Circuit Breaker Tokyo Univ. Sophia Univ. Univ. of Wuppertal ONERA Warsaw Univ. Cracow Tech. Univ. Rolls-Royce Ruston-Homsby imperial College Power Jets Ltd R-R Allison Univ. of Florida Swissauto MSU IUPUI NASA NPS MSNW GPC ETH BBC 1940 1950 1960 1 970 1980 1990 2000 Figure 24: Historical perspective of wave rotor technology. Red: gas turbine application, ( Il'ttfl 1 111g, Blue. refrigeration cycle, I( 111-11111 51111.1 l 11 x I . . 9 1 7 . \i . ‘5‘“. ”‘2‘” ::ll1‘1. '2i‘i‘! 3‘: “I“ 47 1.,z,‘ 1.1. Black: general applications CHAPTER 3: WAVE ROTOR THEORY 3.1 Energy Exchange with Waves To understand the principle of wave rotor operation, it is more convenient first to analyze wave processes that occur in a single stationary channel. This is very similar to analyzing gasdynamic processes in shock tubes. Figure 25 schematically shows a shock tube and its corresponding wave diagram (time- space diagram). The wave diagram describes the wave action by tracing the trajectories of the waves and gas interfaces. It is a plot of the wave motion as a graph of time verses space. When the diaphragm which has separated two gases at different pressures i5 ruptured, two types of gasdynamic waves are initiated at the diaphragm location. A shock wave propagates into the low-pressure gas (driven section) and an expansion wave spreads out into the high-pressure gas (driver section). The shock wave increases the pressure and temperature of the driven section rapidly and induces a mass motion behind itself. The interface between the driver and driven gases, called the contact surface, moves with an induced velocity in the same direction as the incident shock. The expansion wave decreases the pressure and temperature of the driver section smoothly. Therefore, across the contact surface the pressure and velocity are preserved while the temperature and entropy change discontinuously. Now, consider a closed tube which is initially sealed on both ends and contains a gas at rest, as shown in Figure 26. The rapid opening of one end of the tube to a higher or lower pressure media can initiate waves in the tube similar to the rupturing of the diaphragm in a shock tube, as it will be described in the following. Two cases are 48 introduced here utilizing the principle of shock tube operation to increase or decrease the pressure and temperature of the gas inside the tube. Before diaphragm High-pressure flow at rest Low-pressure flow at rest opening After diaphragm 4 opening it at? St ,1 4° '1— s 08“ 0° 1 :5 0 ,' {g} l é§ $0 11 Distance 5’. :1 3 8 a. Figure 25: F low in a shock tube after the diaphragm is ruptured Left end wall Right end wall Gas with zero velocity Pressure ‘ Distance Figure 26: Closed tube containing stationary gas 49 0 Generation of a Shock Wave Assume it is aimed to increase the pressure and temperature of the gas trapped in the tube. This can be achieved by the rapid opening of one tube end to a medium with a higher pressure. When the tube opens instantaneously to a high—pressure gas, a shock wave forms and propagates into the channel as shown in Figure 27. This simulates the compression process of the driven section in the shock tube by suddenly removing the diaphragm. Time High-pressure gas After left end opens Gas at rest Before lefl end opens ‘ Pressure 7 Distance Figure 27: Left end opening to gas at higher pressure 0 Generation of an Expansion Wave To decrease the pressure and temperature of the trapped gas, one end of the tube is opened to a gas with a lower pressure than the pressure in the tube, as depicted in Figure 28. Rapid opening of one end of the tube results in the formation of a centered expansion fan which propagates into the tube, inducing a gas flow out of the tube. Such a 50 scavenging flow process imitates the expansion process of the driver section in a shock tube after sudden removal of the diaphragm. O .E i- Low-pressure After left 835 end opens G ‘ Before left as at rest end opens I Distance 0 1 ' B 1 U) 8 I- o. ‘ V Figure 28: Lefi end opening to gas at lower pressure In the operation of wave rotors, shock and expansion waves may also be generated in other ways. As an example, consider a case where a gas is flowing through a tube at a steady rate as shown in Figure 29. Suddenly closing one end of the tube can generate either a shock wave or an expansion wave. Figure 30 illustrated how an expansion wave can be generated by rapidly closing the lefi end of the tube. Because the velocity of the gas in contact with the closed end must be zero, an expansion wave propagates into the moving gas, bringing it to rest. Also, the pressure and temperature inside the tube drop depending on the strength of the expansion wave, reducing the flow velocity to zero. Using the same concept, suddenly closing the right end of the tube results in the generation of a shock wave. This is shown in Figure 31. The shock wave propagates in the tube from right to left increasing the pressure and temperature inside the tube. ll 0.) I- :3 U) V) d) I- On Distance Figure 29: Opened tube containing flowing gas 6} Dan . s10 I] Iv E a.,, [.— After left end closes Before left _> Flowing gas —-—> —> end closes a 1 Distance 1 l I Pressure _ V Figure 30: Left end closing to generate an expansion wave 52 ll .é’ ”0* a/ l— 8600‘. W ape \ After left .’ -—> —" __'l \I end clOSCS _ Before left _, _. Flowmg gas —’ I end closes ll Distance 2 3 § 9.. Figure 31: Right end closing to generate a shock wave 3.2 Charging Process in Wave Rotors The above gasdynamic phenomena can be used in an array of rotating shock tubes (wave rotors) to increase the pressure and temperature of a flow by expanding another flow which has a higher pressure and temperature. Several configurations can be introduced for this purpose, but only two of them are briefly discussed here. Figure 32 shows a developed (unwrapped) view of the charging process of a wave rotor. This figure shows a sequence of events occurring within the channels moving in the upward direction. Such a representation is called a wave diagram. The wave process occurring inside the wave rotor channels is customarily illustrated by the wave diagram, where the circular motion of the rotor channels is replaced by a straight translatory motion. It describes the rotor internal operation by tracing the trajectories of the waves and gas interfaces. The wave diagram is very useful for visualizing the wave process occurring inside the channels and also for determining wave rotor design parameters, i.e., 53 port opening and closing times and their locations. The utility of the wave diagram is analogous to that of a velocity diagram for a conventional turbine or compressor. In Figure 32, the ducts are set at the correct angle, so that in the rotor reference frame the flow can enter and leave the rotor aligned with its axis. The shock wave trajectory is shown by a dashed line. The trajectory of the interface line is indicated by a dotted line. t Figure 32: Wave diagram of charging process with two primary shock waves The process begins in the bottom part of Figure 32, where the channel is closed at both ends and contains low-temperature and low-pressure flow (e.g., air). When the inlet port opens, the rotor channels are exposed to a high-pressure and high-temperature gas arrived from a heat source (e.g., the combustion chamber). This hot driver gas penetrates the channel. Because its pressure is higher than the pressure in the channel, a shock wave 54 is triggered starting from the lower comer of the inlet port. The shock wave runs through the channel and causes an abrupt rise of local pressure inside the channel. The shock wave speed is higher than the local speed of sound. Within the relevant design space the flow speeds are everywhere subsonic. Therefore, the air/gas interface follows the shock wave with an induced velocity less than the speed of sound. However, behind the shock wave the compressed air has the same local pressure and speed as the inlet driver gas. As the shock wave reaches the end of the channel, the outlet port opens through which the compressed air is then pushed out. At this moment, both the gas and the compressed air column in the channel have the same local pressure and move uniformly with the same induced velocity toward the outlet port. By closing the outlet port, a second primary shock wave originates from the upper outlet edge and propagates from the right to the left. It reduces the velocity of the compressed air to zero to satisfy the zero-slip boundary condition at the end wall. The process is finished when the gas inlet port closes. The moment is timed with the arrival of the secondary shock wave to the upper comer of the inlet port. While the charging process described above is theoretically feasible, it is rarely used in wave rotor designs. Even though the second primary shock wave compresses the air further, the doubly compressed air is not delivered to the outlet port. This may lead to overheating of the channels. Hence, the second primary shock wave has been only generated to stop the compressed air. Therefore, another type of charging processes utilizing an expansion wave is often preferred. This configuration is shown in Figure 33 where the flow is stopped by closing the gas inlet port. This way, an expansion wave originates from the upper comer of the gas inlet port and propagates toward the right end 55 of the channel. Expansion waves (fans) are depicted by thin solid lines. Since this expansion wave induces a flow velocity equal but opposite to that of the gas flow, the gas flow behind this expansion wave is stopped and its local pressure is decreased. Both local and total pressures behind the expansion wave are still considerably higher than the pressure of the fresh air in the beginning of the process. The speed of the expansion wave with reference to the channel wall is equal to the gas speed plus the local speed of sound. As a result, the expansion wave is traveling faster than the air/gas interface, overtaking the air/gas interface before the latter reaches the right end of the channel. The expansion wave reaches the end of the channel at the moment when the upper edge of the outlet port closes the channel. At this moment, the trapped flow within the channel consists of a large part of the hot gas and a plug of compressed air preventing the hot gas to contact the end wall. Thus, during the high-pressure process, penetration of the hot gas into the outlet flow is avoided. Figure 33: Wave diagram of charging process using single shock wave and expansion wave 56 3.3 Principles of Four-Port Wave Rotor Operation A variety of wave rotor configurations have been developed for different applications. The number and azimuthal location of the wave rotor ports along with heat addition schemes distinguish them for different purposes. As described in the previous chapter, four-port configurations have been mainly used as superchargers for internal combustion engines. Three- port wave rotors have been employed in pressure dividers and pressure equalizers in which the pressures of different fluids are increased or reduced. Two-port, four-port, five-port, and nine-port wave rotors have been extensively investigated for gas turbine engine topping applications. As an application of current interest, a four-port wave rotor integrated into a gas turbine cycle is briefly discussed below to illustrate wave rotor operation and options. 3.3.1 Through-F low versus Reverse-Flow Wave Rotors In the described wave rotor illustrated in Figure 6, both gas and air inlet ports are located on one side of the rotor while the outlet ports are located on the other side of the rotor. This configuration is known as the through-flow (TF) wave rotor in the literature. Alternatively, another type of wave rotor has been designed where the fresh air enters and exits at the same end of the rotor (air casing) while the burned gas enters and exits the rotor at the other end (gas casing). This configuration is called reverse-flow (RF) wave rotor as shown in Figure 34. These two configurations may provide identical topping and overall performance enhancement, but they differ substantially in their internal processes. 57 Combustion Chamber Wave Rotor Compressor Turbine Figure 34: Schematic of a gas turbine topped by a reverse—flow four-port wave rotor In a TF four-port wave rotor, both hot gas and relatively cold air traverse the full length of the rotor, keeping the wall at a relatively uniform intermediate temperature. This self-cooling feature of TF wave rotors has prompted interest in them for gas turbine engine topping applications where gas temperatures are high. The RF configuration does not inherently result in such a self-cooled rotor. The cold air never reaches the other end of the rotor as seen from Figure 34. As a result, the air side of the rotor is relatively cool while the gas side of the rotor is relatively hot. To achieve a better self-cooled RF design, a mirrored reverse—flow cycle design of the RF configuration can be constructed, which orients the cycle alternately right and left on the rotor [189]. This approach introduces symmetry and assures that both sides of the rotor are washed by the relatively cold fresh air. Unfortunately, it also poses severe ducting challenges. However, for small gas turbines the gas temperatures are ofien lower than those of large scale gas turbines. Therefore, the RF approach seems to be the viable choice for microturbines. Also, it has been claimed that the RF cycle provides better separation of air and gas in the rotor channels [2]. Due to this separation of air and gas regions in the channels, the analysis of the fluid flow inside RF wave rotor channels is easier. Knowing about all these facts, RF 58 configurations have been mostly used in the relatively low-temperature application of car engine supercharging although such configuration for gas turbines has been also investigated [131, 189, 198]. The General Electric Company has obtained experimental data on a gas turbine engine enhanced by a RF wave rotor [126]. 3.3.2 How Does it Work Inside? Figure 35 represents NASA wave diagrams [168] for through-flow (left) and reverse- flow (right) four-port wave rotors for one cycle operation of the rotor. The journey of a channel of the wave rotor is periodic. The top of each wave diagram is considered to be looped around and joined to the bottom of the diagram. This requirement presents a fundamental challenge in the simulation and design of wave rotors. A successful prediction of the wave rotor implies that the state of the working fluid in the channel at the end of the cycle must be the same as that postulated at the beginning of the cycle. To show how a four-port wave rotor works, the events occurring in one cycle are now described. For the RF configuration, the process begins in the bottom part of the right wave diagram where the flow within the channel consists of a large part of the hot gas and a buffer layer separated by a contact surface. For the TF wave rotors (lefi), the gas fills the whole channel. As the right end of the channel opens to the relatively low- pressure outlet port, an expansion fan originates from the leading edge of the outlet port and propagates into the channel, discharging only the gas to the turbine. The expansion fan reflects off the left wall and reduces the total pressure and temperature in the channel further. This draws fresh air provided by the compressor into the channel when the air inlet port opens. When the reflected expansion fan reaches the outlet port, it slows the outflow and reflects back as compression waves, while the outlet port closes and halts the flow inside the channel. The compression waves form a single shock wave as they travel 59 toward the inlet port. As the shock wave reaches the upper corner of the inlet port, it closes gradually. At this moment, the channel fluid is at rest relative to the rotor. 60 :6: Mom atom 5wa 582 o>m>> tonnfifl Camry Bocémbfiz can an: Bocéwsohfi H8 mEBwEu “335 ”mm Eswi S x‘l m tom 339:3 2:32; 33 0:. h: 0 .n h _ s E Ex. E5 3339:. 2.0.. .ommoEEoo 83% ton .95 9530:. :9: 50:530 thu ton .9558 2:395 :9: 59:5 ch :ozflom « x .Al o E 333.. 3:230: n 50:.- toa "9558 2:305 30.. an 0:. .5 O .n a P E can 3.5 2:32.. 33 3395200 Soak tom “2:. 3. Banana :2: tom .9353 me .0539 50.5 2339:. :9: hop—=5 OF 61 The above sequence of events is called the low-pressure part of the cycle (scavenging process). Its purpose is to deliver a high-pressure gas into the turbine, partially purge the rotor channels, and ingest fresh air received from the compressor. In the high-pressure part of the cycle (charging process) that follows, the rotor channels are exposed to the burned gas arrived from the combustion chamber. The wave phenomena here are same as those described in Figure 33. The only difference here is the existence of a reflected shock wave in the NASA wave diagram. In fact, as the shock wave reaches the end of the channel and the outlet port opens a reflected shock wave originates at the lower outlet edge, propagating back into the channel. The reflected shock wave compensates for the combustor pressure loss. On the other hand, for an ideal combustion chamber (no pressure loss), the reflected shock wave would not appear. This has been shown in Figure 77 which is discussed later. The double-compressed flow behind the reflected shock wave leaves the wave rotor toward the combustion chamber. In the RF configuration the discharged flow into the burner is pure air, while in the TF configuration usually both air and once-burned gas are delivered into the burner. Detailed fluid flow investigations have suggested that approximately 30 to 50% of burned gas is recirculated to the combustion chamber in the TF configuration [180]. Again, a favorite case is considered when the closure of the gas inlet port is timed with the anival of the reflected shock wave. At this moment, an expansion fan originates from the upper corner of the inlet port and propagates toward the other end of the channel which eventually brings the channel flow to rest. When the expansion fan reaches the end of the channel, the outlet port closes and the flow in the rotor channels stops. At this point, the flow with zero velocity is at nearly 62 the peak pressure and temperature of the cycle. It is now ready to be discharged into the turbine during the low-pressure process. 3.4 Gasdynamic Equations To find flow properties inside wave rotor channels, it is required to predict the gasdynamic processes occurring inside the channels using boundary conditions provided by a thermodynamic cycle analysis. Proper design of a wave rotor requires a reasonable solution for the internal flow field. To derive necessary equations for a gasdynamic analysis, several assumptions are made throughout this work. The aspect ratio of the wave rotor channels is assumed to be large enough, so the flow can be treated as one-dimensional. The flow within the rotor is considered frictionless and adiabatic. However, the wave rotor efficiency is used to account for dissipation losses inside the channels. The gases are treated as ideal gases. 3.4.1 Moving N ormal Shock Wave Relations Consider a moving normal shock wave which propagates with an absolute velocity ws from right to left into a channel which contains a flow moving with velocity u; from left to right, as shown in the top part of Figure 36. The flow velocity decreases from u, to u; due to the mass motion induced by the shock. The term unsteady reference frame is attributed to this case. To find the flow properties in region 2 (compressed flow after the shock), the governing normal shock equations which are only valid for a stationary frame of reference can be used. In a reference frame moving with the shock wave, the shock appears stationary and the downstream and upstream flow velocities are u, + ws and u; + ws , respectively as shown in the bottom part of Figure 36. 63 Movmg shock wave ——> _.__> u, + wS E 112+ Ws Stationary shock wave —> —p Figure 36: Transformation from a moving shock wave to a stationary shock wave Applied to a control volume considered around the stationary shock wave, continuity, momentum, and energy equations become: ,0,(W5 +ul):p2 (W5 +u2) (2) pl‘l‘p/(WS‘H’IY =p2+p2(w5+u_,)2 (3) 2 2 h,+(ws+u') :h2+(ws+u2) 4 2 2 () Assuming a calorically perfect fluid, solving the above equations leads to the relations given below, which give the local temperature ratio (T 2 /T 1) and local density ratio (pg/p1) across the shock wave as a function of the local pressure ratio [75 =p2 /p, : T y+i+ 5 -4=Hsy’ (a T: I y+1 +— s y-I “Ii-(17.) 3i: 7’1 (o pl XII—{+175 7-1 64 As seen, local thermodynamic properties across a moving normal shock are physically independent of the flow field velocities. Furthermore, the induced velocity can be expressed as: l _2_7_ + 1 uimluce—shuck = 5:11. (”S _ I) y 7 '— I I] _ V S + 7 +1 (7) where 0, represent the speed of sound in the undisturbed region 1. Therefore, the flow velocity after the moving shock in region 2 becomes: u 2 = “I — u induce—shock (8) The moving shock velocity can be obtained as [216]: 1 w, = a, J%(HS - 1)+ 1 (9) However, the velocity of the shock relative to the gas in the region 1 becomes ws - u 1 or: wS =a, %l(175—1)+1—u, (10) Furthermore, the shock Mach number is introduced as: ML: fil<fls_1)+1 (1]) M = S a, 2}! Finally, the entropy change across the shock is obtained by: A; = CPln? —C,,1np-’ l l (12) It is worthwhile to note that the total pressure increases across a moving shock while it decreases across a stationary shock wave. Also, the total enthalpy is not constant across 65 a moving shock wave. It increases for a moving shock wave, but it remains unchanged for a stationary shock wave. 3.4.2 Expansion Wave Relations Due to the isentropic nature of expansion waves, the method of characteristics can be used to find the flow properties across an expansion wave. This method is a very general and powerful technique for analyzing compressible flow. For the specific problem considered in this study, only one-dimensional version of this method is used. Because the method of characteristics is a well-know theory, only the application of this method is used here and details of this theory can be found in the literature [216]. Consider any given point (x, , t,) in the x-t plane as shown in Figure 37. It is possible to find a specific path through point (x,, 1,) along which the following equation holds: a =c0nstant=J+ (13) y—l ll+ where u is the flow velocity and a is the sound speed. Such a path is called a C+ characteristic line (right-running wave) in the x-t plane. This specific path is chosen so that its slope becomes dt /dx=1/(u + a). In addition, a C_ characteristic line can be found through the point (x, , 11) in Figure 37, where the slope of the C__ characteristic (left-running wave) is dt/dx=1/(u - a) and along which the following equation obtains: a =c0nstant=J_ (14) 7—1 "— The constant values of J+ and J_ are known as the Riemann invariants. 66 dx/dt=u—a dx/dt=u+a ‘ZJ § 3» e g .9 C.) e (U a? ‘A 0 ‘ V xl x Figure 37: Illustration of the characteristic lines in the x-t plane With the above results, it is possible to solve for the flow field in a one-dimensional expansion wave. As an example, consider again the generation of an expansion wave by suddenly closing a tube, as previously described in Figure 30. This figure is modified in Figure 38 to include a C_ characteristic line as well. Note that here the C + characteristic lines are physically the paths of the expansion waves in the x-t plane. 11 15‘, 11.30812)” 6’6 (C a 11'” es) Figure 38: Left end closing to generate an expansion wave 67 Knowing the flow properties in front of the head of the expansion wave (region 1), solution for the flow behind the tail of the expansion wave (region 2) can be obtained as now described. Because J_ is constant through the C_ characteristic line, Eq. (14) applied to both regions 1 and 2 gives: 2 2 , u,— a, =u — a‘ =J_ (15) 7—1 2 7—1 However, the wall condition on the lefi side implies that u2=0. Therefore, the sound speed ratio between region 1 and 2 can be written as: 3.2. -_- 1-2/:1 u_, (16) a, 2 a, or, 2 T _ _Z_=I:1_7_Iu_’] (17) T, 2 0, Because the expansion flow is isentropic, therefore: _7_ ~’_7 -1 _ —-l p? =[—T—~’—Jr =[1——7 I 59—]7 (18) PI T1 2 a] and, _’_ .3. ._l _ -/ pg: 11.], =1__}’_{31_,_7 (19) pl TI 2 at The expansion wave induces a mass motion in the direction opposite to its propagation, but equal in value to the flow in region 1. Therefore, it is also possible to find a relation 68 for the induced velocity based on the pressure ratio across the expansion wave. This can be easily found by using Eq. (18) as: 2 . 2r “induce—exp umiun = a, I - (L) (20) 7‘1 F! It can be shown that this relation obtained for the induced velocity is independent of flow velocities of the medium in which the expansion wave moves. Here, the velocity of the flow in region 2 was zero, however, Eq. (20) always calculates the change of flow velocity due to the wave expansion. This conclusion is also valid for the induced velocity generated by a moving shock wave described before, as shown in Eq. (7). Both induced velocities are only functions of the pressure ratio across the waves, the speed of sound, and the 7 value of the downstream flow. Finally, it is important to mention that the head of the expansion wave moves with the velocity u, + a, to the right because it is a C + characteristic line. The same reasoning shows that the tail of the expansion wave moves at the slower velocity 0 + a; in the same direction. Hence, the expansion wave spreads out as it propagates down the tube. 69 CHAPTER 4: THERMODYNAMIC ANALYSIS 4.1 Microturbines A growing market for distributed power generation and propulsion of small vehicles has motivated a strong interest in design of small gas turbine systems in the range of 30- 300 kW. Known as microturbines, they are now widely used in the US for distributed power generation, shaving peak loads, and providing backup power for critical needs. They propel small commercial aircraft, unmanned air vehicles (UAV), and terrestrial vehicles. Microturbines are often the preferred alternative to IC engines, due to their higher power density and robustness. They present several advantageous features such as compact size, simple operability, ease of installation, low maintenance, fuel flexibility, and low NOx emissions. Furthermore, due to the recent electric-power crises and environmental concerns, a strong interest in the research, development, and application of microturbines has been stimulated Despite their attractive features, compared with larger gas turbines, microturbines suffer from lower thermal efficiency and their relative output power, due to their limited cycle pressure ratio and peak cycle temperature. For many applications improvement of their performance is desirable to enhance advantages over competing technologies. To achieve such improvements, current efforts are mainly focused on utilizing heat recovery devices and developing new high-strength high-temperature materials for turbine blades [217]. Geometries of microturbines make blade cooling very difficult. Hence, their lifetimes using typical materials used for larger gas turbines are shorter [218]. Therefore, there is significant research toward developing advanced metallic alloys and ceramics for high-therrnal-resistance turbine wheels used in microturbines[219, 220]. 70 Currently, recuperators play a key role in performance enhancement of microturbines. For example, experimental and theoretical research has shown that microturbines with pressure ratios of 3 to 5 without recuperation systems achieve only about 15 to 20% efficiency [221, 222]. Utilizing conventional recuperators based on the use of existing materials can improve the thermal efficiency of microturbines up to 30% [222-225]. An excellent example of a commercial microturbine with a recuperator is the Capstone 30 kW unit, with an efficiency of 26% using natural gas fuel [226]. Despite the attractive feature of the recuperator concept, a recuperator introduces pressure losses reducing the output power and it adds about 25 to 30% to the overall engine manufacturing cost, which is a challenge for commercialization of microturbines [227-229]. The current trend of the microturbine market is to reduce the investment cost per kW. Therefore, alternative devices need to be considered to achieve higher performance at lower component costs. Topping a microturbine with a wave rotor device is an appropriate solution. Wave rotor investigations [154, 164, 198-200] have shown a significant potential for performance gain in smaller gas turbines, where the compressor pressure ratios are typically lower than those of larger machines. The objective of the present chapter is a comprehensive and systematic performance analysis of two actual microturbines known as the C-30 and C-60 engines made by Capstone Turbine Corporation which are topped with a four-port wave rotor in various wave-rotor-topping cycles. The challenges and advantages associated with the different implementation cases are discussed. While the performance evaluation of several gas turbine engines has been studied extensively [154, 198-, 200], to the knowledge of the author, there exits no comprehensive work investigating the potential benefits of various 71 implementation cases of wave rotor topping cycles for small gas turbines. The presented results have been obtained using basic thermodynamic equations along with the wave- rotor characteristic equation previously validated using computational tools [154]. The model can be employed to predict the performance improvement of various wave-rotor- t0pping cycles without the need for knowing the details of the complex fluid mechanics within the wave rotor. 2 4.2 Gas Turbine without Recuperation 4.2.1 Implementation Cases There are several possibilities to top a gas turbine with a wave rotor. Considering possible design restrictions and preferences, five different advantageous implementation cases for a wave rotor into a given baseline engine can be introduced as follows: Case A: same compressor, same turbine inlet temperature Case B: same overall pressure ratio, same turbine inlet temperature Case C: same combustor Case D: same turbine Case E: same compressor, same combustion end temperature These five cases have been described in detail in the following. According to the state numbering introduced in Figure 4 (or Figure 34), Figure 39 visualizes all five cases in schematic T-s diagrams. Path O-lb-4b-5b represents the baseline cycle and O-li-Zi-3i-4i-5i (i=A, B, C, D, E) indicates the wave-rotor~topped cycles, where the subscripts indicate the case. One of the five cases might be preferable for a practical design. However intermediate design cases are possible. 2 Parts of materials presented in this chapter have been accepted for publication in 2005 ASME Journal of Engineering for Gas Turbines and Power. 72 o-lb-4b-5b baseline engine o'lrzr3r4i-5i topped engine i=A, B, ..., E ,3A _ Tiara»- _ _ Temperature Entropy Figure 39 : Schematic T-s diagrams for a baseline cycle and five wave-rotor-topped cycles Case A: In Case A the pressure ratio of the compressor is kept unchanged, so the physical compressor of the baseline engine can also be used for the wave-rotor-enhanced engine provided the mass flow is kept approximately the same. The pressure in the combustion chamber of the enhanced engine is increased by the compression ratio of the wave rotor. This may require modifications to the structure of the combustion chamber and to the fuel injection system. The heat addition in the combustor is the same as for the baseline engine, but it takes place after the energy exchange in the wave rotor, hence the heat addition starts at a higher temperature. Thus, the combustion end temperature is even higher than that of the baseline engine, possibly requiring additionally a thermal enhancement of the combustor structure. The turbine of the topped engine might need to be adapted to efficiently utilize the higher pressure ratio. The turbine inlet temperature, 73 however, is the same as that of the baseline engine. As will be shown later, this implementation case provides the highest thermal efficiency and specific work and the lowest value of SF C . Case B: In Case B the overall pressure ratio for the wave-rotor-enhanced engine is kept equal to that of the baseline engine, so that the combustor works under the same pressure. However, for the wave-rotor-topped engine, the heat addition in the combustor and the combustion end temperature are greater than those of the baseline engine. This may require some adaptation of the combustor, especially in the outlet region. The turbine and compressor work with lower pressure ratios, reducing the design challenges. Thus, both may be adapted advantageously. This might reduce the cost of the compressor and turbine due to reduction of stages in multi-stage types (mostly axial), or due to reduction of the tip diameter in radial types (mostly single-stage). With a smaller tip diameter the wheels can be manufactured more economically over a shorter time from cheaper materials with less strength and on smaller machines. Besides an attractive performance enhancement, this case additionally provides the highest turbine outlet temperature of all five cases investigated. The temperature of the leaving exhaust gas is much higher than that of the baseline engine. Therefore, this case is attractive for an external heat recovery application or for internal recuperation that can enhance the performance further. Case C: Case C assumes that it is desirable for the wave-rotor-enhanced engine to use the unmodified combustor of the baseline engine. So the overall pressure ratio and combustor inlet and outlet temperatures for the wave-rotor—enhanced engine are kept equal to those of the baseline engine. The heat addition in the combustor is consequently 74 the same.3 The implementation of the wave rotor considerably reduces the pressure ratio of the turbine and compressor. The compressor pressure ratio is as low as in Case B, and the turbine pressure ratio and turbine inlet temperature are even lower than those in Case B. Thus, the turbine and compressor could be made from less thermally resistant material. Compared to the baseline engine, they also could be smaller and hence less expensive, as discussed in Case B. This might be the main implementation reason because unfortunately, but not surprisingly, the unambitious combustor constrains the performance enhancement. It is nearly negligible for the smaller C-3O engine and even negative for the C—60 engine. Case D: Case D employs the same physical turbine as the baseline engine. Due to the wave-rotor—topping, the compressor needs to produce a lower pressure ratio than that of the baseline engine. This allows for a smaller and less expensive compressor as discussed for Cases B and C. The pressure in the combustion chamber and the combustion end temperature are higher than those of the baseline engine, but lower than those of Case A. Hence, less effort might be required to adapt the structure and fuel injection of the combustion chamber. As a result of the lower pressure ratio in the compressor, hence lower compressor discharge temperature, the heat addition in the combustor has to be more than that for the baseline engine to utilize the same allowed turbine inlet temperature. This case gives the second highest performance increase for both baseline engines. Case E: Case E is similar to Case A but the combustion end temperature (the cycle peak temperature) is restricted to the turbine inlet temperature of the baseline engine in 3 The wave rotor compression efficiency is greater than the compressor efficiency. Therefore, the combustor inlet temperature is in fact negligibly smaller and hence the heat addition is negligibly greater than that in the baseline engine. 75 order to avoid additional thermal requirements on the combustor design. The overall pressure ratio is the same as in Case A because this case employs the same physical compressor as for the baseline engine. Thus, the overall pressure ratio is greater than that of the baseline engine, by the wave rotor pressure ratio. The heat addition in the combustor is less than that for the baseline engine because to the wave rotor compression work is added to the fluid before combustion. The turbine in the topped cycle works with a slightly greater pressure ratio than the turbine of the baseline engine, but the turbine inlet temperature is less than that for the baseline engine. In fact, it is the lowest of all cases investigated. This may give the option to produce the turbine wheel at a lower cost out of less thermally resistant material. 4.2.2 Thermodynamic Calculations To evaluate the performance enhancement of topping gas turbines with wave rotors, a computer program based on a thermodynamic model has been developed to determine the thermodynamic properties of the gases in different states of the cycles. The results are used to calculate the theoretical performance (expressed by thermal efficiency ,7] , net specific work w , and specific fuel consumption SF C) and the actual T-s diagrams of both wave-rotor-topped and baseline engines. The methodology is similar to that introduced by Wilson and Paxson [154] with some modifications. The component performance parameters for the baseline engines are based on information provided by the manufacturer and are listed in Table l. 76 Table 1: Baseline engine data, assuming T 0 =300 K, Cp,,,—,=l.005 kJ/kgK, Cpg..s=l.l48 kJ/kgK, yair=l.4 , yga, =l.33 Baseline engine C-30 C-60 turbine inlet temperature Tm, (11151566? [1,: $272576? I]; 466.5 K 505.4 K compressor outlet temperature Tm, :3 80° F) (450. F) compressor pressure ratio p, 1M0 3.6 4.8 compressor isentropic efficiency rzc 79.6% 82.6% turbine isentropic efficiency Ur 84% 85% compressor polytgpic efficiency mpg 82.9% 85.9% turbine polytropic efficiem ”N 81.7% 82.3% combustor pressure ratio [Lo—”d, 0.98 0.98 For each engine it is assumed that the compressor inlet condition is known and is the same for both baseline and wave-rotor-enhanced engines. Considering the same “aerodynamic quality” of the wheels, the polytropic efficiencies are kept the same for the enhanced and baseline engine, for the compressor and turbine respectively. Incomplete combustion of the fuel is reflected by a combustor efficiency of 98% (17g = 0.98). No pressure losses in the intake air filter, exhaust silencer and additional piping, or heat losses or mechanical losses are taken into account. Such losses reduce the predicted performance. The gases are treated as ideal gases. For both air and burned gas constant values of specific heat coefficients (Cpair = 1.005 kJ/kgK, Cpgas = 1.148 kJ/kgK) and specific heat ratios (nu-r: 1.4 , ygas= 1.33) are considered. This assumption simplifies the performance calculations significantly without affecting the qualitative nature of the results. Air enters the compressor at 300 K. In the following it is assessed how the wave-rotor-topping enhances the performance of C-30 and C-60 baseline gas turbine engines. 77 0 Path 0-1: Compressor With the given compressor inlet temperature (T 0=T ,0), the compressor outlet total temperature and pressure are calculated from the adiabatic relations: T run-I mam—1p. -1) 77C :21 = EL = ”C (22) p10 p0 where the compressor isentropic efficiency (tic) relates the compressor pressure ratio (17 C) to the compressor polytropic efficiency (npc) through: run-1 17 —1 77c = C r (23) yr"! —1 17C Ymr’IPC‘ _1 For Cases A and E, the compressor pressure ratio is equal to that of the baseline engine (e.g. for the 030 engine 17C = 3.6 and for the C-60 engine 17C = 4.8). However, for Cases B and C its value is calculated by dividing the baseline compressor pressure ratio with the wave rotor compression ratio (PRW = p,2/p,,). The wave rotor compression ratio plays an important role in the performance analysis of wave-rotor—tapping engines by significantly altering the engine performance. A higher value of a PRW leads to a higher engine performance. F atsis and Ribaud have varied PRW from 1.4 to 3 in their calculations [198, 199]. PRW values lower than 1.8 have considered unacceptable due to the poor compression processes inside the rotor. The authors believe that a PRW value higher than 3 is unrealistic. This is confirmed by detailed calculations in a four-port wave rotor by Welch [164]. Wilson and Paxson have used PRW = 1.8 from GE data [154]. They have used the term “advanced” for their wave rotor working with PRW = 3.6. Therefore, a 78 wave rotor compression ratio of 1.8 appears to be conceivable for the envisioned application and is chosen for the following discussion. In this work all performance plots are shown for various wave rotor pressure ratios, indicating its effect on cycle performance enhancement. For Case E, the value of Hg is calculated in a way that keeps the pressure at the turbine inlet equal for both baseline and topped engines. This calculation can be performed by inreversly solving of Eq. (21) through (41). Finally, the compressor specific work is obtained as: C . T! r..,,-1 WC = Cp... (T., —T..) = fume —1) (24) 77C 0 Path 1-2: Compression in the Wave Rotor The flow properties after the wave rotor compression process are obtained from the adiabatic relations similarly to those of the compressor: T T,_, = T,, + ——"— PRW - 1 (25) ”WC ELL-zfifliszw 'Hc (26) p10 pr] p10 where 27% is the wave rotor compression efficiency. Later my; , which is defined as the wave rotor expansion efficiency, will be used. Consistently with previous wave rotor investigations [154, 164, 198, 199], the wave rotor compression and expansion efficiencies are assumed as mm = my; =0.83 in this study. In fact, the efficiencies are not constant values. However, the range of variations have been almost determined. In previous wave rotor studies it was not possible to evaluate mm and my}; separately. Instead, their product was determined. Moritz found mm - :ng = 0.6 in experiments at Rolls-Royce [130]. Reference [153] has reported a range of 0.77 < mm = 17mg < 0.86. 79 Wilson and Paxson assumed mrc = ”WE = 0.83 based on previous work on wave rotors [154]. Recently, Fatsis and Ribaud have taken into account the effect of wave rotor efficiencies on the performance of several different types of gas turbine engines topped by a wave rotor [198, 199]. They have carried out their calculations with efficiencies in a range between 0.8 < mm = my}; < 0.86. For a small size turbojet engine considered in their study, they have shown that the gains of specific power and specific fuel consumption are sensitive to the values of wave rotor efficiencies. These authors have emphasized that it is important to optimize efficiencies for the compression and expansion processes. 0 Path 2-3: Combustion Chamber The value of firel/air ratio (f =mf/mair) can be obtained by applying the energy (first law) equation to the combustion chamber: ”Q f hPR =(1 + f)Cpgas 7:3 _ Cpair T12 (27) where hpR = 43000 kJ/kg is the heating value used for all calculations here. This equation gives: __ Cpgas 7‘13 —Cpair 22 ”Q hPR — Cpgas 7:3 (28) Alternatively, f can be expressed based on the turbine total inlet temperature (T ,4) and the compressor total exit temperature (T ,1). For this purpose, the wave rotor compression and expansion specific work (per unit mass of air flow) are defined respectively as follows: WWC = Cpair (Tu _ TH) (29) wwg = (1 + f ) C19,... (Tu - 7.4) (30) 80 Here, it is considered that m, = n'z, = m and m, = m, = mg, + #2,. Other cycles exist in which the mass flow rates m, and m, are not the same, and, correspondingly the mass flow rates m, and n'z, may not be equal either. These cycles are not considered here. Using Eqs. (29) and (30) , Eq. (27) can be expressed as: 779 f hm = (1+f) Cpg... T... + WW5 - wwc -Cp...-, TH (31) Because the net output work of the wave rotor is zero (wwc = WW5), solving for f leads to: _ Cpgas 7:4 —Cpair 7;] ”Q hPR -Cpgas 7:4 f (32) For Cases C and E, T ,3 is equal to the baseline inlet turbine temperature, therefore Eq. (28) is used to calculate f. For Cases A, B, and D, where T,., is a known value and is equal to the baseline inlet turbine temperature, f can be obtained from Eq. (32). The relation between T,3 and T ,4 can be found by equating Eqs. (29) and (30): T mr‘", C . 7:3 :734 +l: tl [PRWyr‘m —1]] pair (33) 77wc (1 + f) C10,... Finally, the total pressure after the combustion chamber is obtained: & : 111$ = 17 P10 Pr) on PR,,, .17, (34) comb 0 Path 3-4: Expansion in the Wave Rotor To obtain the turbine inlet total pressure (19,4), it is convenient to express the wave rotor compression work and the expansion work in terms of pressure ratios: ch=m Cp..,(T» -T,) = ”” PRW'ZT—z ”we ' C .T run-l m pair tl( ) (35) 81 -I. 'yl‘l‘ P0 :(muir +m /)Cpgas (773—14): (mair +m /)Cpgus ”WE 7:3 I—£_—_u—] (36) 7mm 17 PR comb Where P0=p,4 /p,, is the pressure gain ratio across the wave rotor. Equating the compression work to the expansion work leads to: 7 gm ‘ I " ng 17 PR comb C . T nr’ —’-’”—”-—"(PRW —1]= (I+f)Cp... n... T” I— {—Ig—fl <37) ”WC Substituting 7),; from Eq. (33) into Eq. (37) and some algebra gives: r ‘ r,.a,/rg...-l A—1———B P0: 17m, PR 1- '7” "I” > (38) 1+A——B ( ”WC , where C . : pal, (39) (1+f)Cpg... B_Tz___1_ [PRW (r... -l)/r.,,-, _1] (40) Tu Equation (38) is a modified version of the “wave-rotor characteristic” equation introduced in the literature [154]. This equation represents the performance of the wave rotor. It predicts the pressure gain ratio across the wave rotor as a function of the temperature ratio across it (7)., /T,1). For ”WC = my); = 0.83 , f = 0, 1760”,), = 100%, and PRW= 1.8, Figure 40 clearly indicates monotonically increasing of the pressure gain as a function of the temperature ratio across the wave rotor. For a fixed turbine inlet temperature (e.g., 7",.) = constant), using a smaller compressor with a low pressure ratio leads to a higher pressure gain across the wave rotor. Similarly, for a fixed compressor 82 pressure ratio, increasing the turbine inlet temperature results in a greater pressure gain across the wave rotor. The most significant performance gain has been found for engines with low compressor pressure ratios and high turbine inlet temperatures [154, 164, 198, 199] Equation (38) can be also used again to investigate the influence of the wave rotor compression ratio on the pressure gain ratio across the wave rotor. This is shown in Figure 41 for ”WC: 17m; = 0.83 ,f= 0, 17mm), = 100%, and 7)., /T,, = 2.2. The plot clearly indicates that the pressure gain ratio across the wave rotor is an increasing function of the wave rotor compression ratio, while the highest relative gains are obtained at low compressor ratios. Therefore, higher values for PRW results in a higher net absolute performance gain. 1.2 as: .- 11.: f g 1.15 I a m s a r- o 1.1 g I 3 . 2 E ,. O s 1.05 m a > a L- 3 b.5411.8112.11I2.4112.71 T3 Wave Rotor Temperature Ratio (run ,1) Figure 40: Pressure gain ratio across the wave rotor versus temperature ratio across it 83 1.2 0::- / :3 / ‘35 115 // N m _ C a r o / 2 1.1 3 .. 1 / m 2 / a, _ O *5 1.05 K 0 > N 3 11 1.5 2 2.5 T3 ‘ 13.5 Wave Rotor Compression Ratio (PRW) Figure 41: Pressure gain ratio across the wave rotor versus wave rotor compression ratio By using Eq. (3 8), the turbine inlet total pressure is obtained as: _p_IL=£1_I_)L/_=P0.17C , (41) P10 Pu pro For Cases C and E, where T ,4 is an unknown value, the turbine inlet temperature is obtained by using Eq. (33), and T,3 is equal to the baseline inlet turbine temperature. 0 Path 4-5: Turbine The turbine specific work (per unit mass of air flow) can be calculated knowing the pressure ratio across the turbine: has -1 w, = (1+ f)Cp.... (TM —T..) = (1+ m. 6p... 7,. I—(fl) <42) :4 assuming the turbine expands the hot gas leaving the wave rotor to atmospheric pressure (p,5=p0). The isentropic turbine efficiency (m) can be obtained from the polytropic turbine efficiency mar: 84 ( p0 )(Ym.—l)’7rr ’ll’g... _ I ’77- : pl4 (43) (&)(y£m_l)/7rm _1 p14 Therefore, the total temperature of the gas leaving the turbine can be calculated as: W T =T — T 44 ’5 ’4 (I+f)Cpgm ( ) After calculating the thermodynamic properties of all states in the cycle, it is possible to calculate the engine performance parameters. The net specific output work produced by the engine can be calculated by subtracting the turbine specific work from the compressor work: 71:1“ —I 1 Cpuir 7; 7017‘] W=77MWT_WC=77M777CpgusTr4 1_(££) 7W _———_0(17C7‘”" _1) (45) :4 77C Where pm is the turbine shaft mechanical efficiency. In this study an ideal transmission case (77M =100%) is considered. With the amount of specific heat addition through the combustion process being defined as q=f. hpR , the thermal efficiency can be written as: rgui —l 1 Cpuir 7; 7"” -l m, m Cp... T... I—(Ei) ’e — ———0(17c — I) W p14 77 C ’7 = — = ' (46) q f hPR Finally, the specific fuel consumption (SF C) is calculated from: SF C = —f— (47) W 85 4.2.3 Predicted Performance Results Cases A and B are the most common cases discussed in the literature, therefore, they have been discussed here in more detail. Case A: Figure 42 shows the actual T-s diagrams for the baseline engines C-30 and C-60 as well as for the topped engines, simulated with a wave rotor pressure ratio of 1.8 (PRW = 1.8). The overall pressure ratio of the enhanced engines is 1.8x3.6=6.48 and l.8x4.8=8.64, respectively. The T-s diagrams qualitatively show that the topped engine has a higher thermal efficiency compared to the baseline engine. This is because the turbine has a higher specific work output, while the consumption of specific work by the compressor and specific heat addition to the cycle remain the same as for the baseline engine. Details of calculations for creating such plots are described step by step in Appendix B. 86 1500 - 045-4195,, baseline engine 0-1 A-Z {3 .1'4 A-S A topped engine 3A 1200 ' Ttnrbine 4A 4 .— _ U _________________ Q h _. _. _. \ E \ a 900 ' \ ‘5 5.. a 5* 3 o. E 600 ' E- lb=lA 300 ' . 0 G30 Engine 0 A l L L j 1.5 1.8 2.1 2.4 2.7 3.0 Entropy lid/kg K] 1500 , 0-lb-4b-5b baseline engine 0'1A'2A'3A-4A'5A topped engine 3A b Tiarbine _______________ _ 1200 ' E i‘.’ 900 ' :1 3 z o. g 600 P l- ]b 300 ' . 0 C-60 Engine 0 l l l l I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy lkJ/kg K] Figure 42: T-s diagrams for the C-30 and C-60 wave-rotor—topped engines for Case A implementation 87 Figure 43 illustrates the increase of cycle thermal efficiency (dash dot) and specific work (dashed), and the decrease of specific fuel consumption (solid) with increasing wave rotor pressure ratio PRW for both the C-30 and C-60 topped engines. The plot visualizes how the effect develops from the baseline engine with PRW =1 until PRW =2 which might be a practical limit for the investigated application. However, if the wave rotor pressure ratio increases beyond this limit, the trend already shows that the rate of increase of the effect diminishes while technical problems may increase. With a conceivable wave rotor pressure ratio of 1.8, the thermal efficiency of the baseline cycle increases from 14.9% to 20.0% for the C-30 engine and from 19.4% to 24.2% for the C- 60 engine. Simultaneously, the specific work increases from 128 kJ/kg to 171 kJ/kg for the C-30 engine and from 184 kJ/kg to 231 kJ/kg for the C-60 engine. The specific fuel consumption (SF C) of the C-30 engine decreases from 0.156 kg/kN.s to 0.116 kg/kN.s and it reduces from 0.120 kg/kN.s to 0.095 kg/kN.s for the C-60 engine. A better picture of the performance improvement is obtained by calculating the relative increases of thermal efficiency, specific work, and the relative decrease of SF C as shown in Figure 44. For Case A, the relative increases of thermal efficiency and specific work (dash dot) are precisely the same as it is obvious fi'om Eq. (46) where the heat addition q=f. hPR is the same for both topped and baseline engines. For a wave rotor pressure ratio of 1.8, Figure 44 indicates an attractive relative performance improvement in thermal efficiency and specific work of about 33.8% and a 25.2% reduction in SF C (solid) for the C-30 engine. The C-60 engine shows a 25.1% enhancement in thermal efficiency and specific work and a 20.1% reduction in SF C. 88 Overall Pressure Ratio — T“=1116.5 K 7; 0.28 5 0.28 “95033 240 z“ ; =0.82 i‘ 0.24 - 0.24 =0.98 .. 220 3’ ~ A V ' 3’ I: 0.2 - 0.2 200 \ .9 - 1 ii I 5 :‘5 g 0.16 7 g; 0.16 180 1; In ‘9 g g . a: o 0.12 - m 0.12 160 51;; 3 ' 3 E — a. u 0.08 — 0.08 140 w 9.: _ g _ n. 0.04 - 0.04 120 ‘0 : L 0 — 0 100 Wave Rotor Compression Ratio (PRw) Overall Pressure Ratio 3 Tu=1227.6 K 13 0.28 _- 11,5006 2' Z =0.82 1‘ 0.24 — =0.9s m r A i . g. g 0.2 - \ a : e 3 ._ E g 0.16 L 2 .g in _ 2 3 s . v: .. o 0.12 - w E _ h o 0 . e - a o 0.08 5 u, SE . g - o. 0.04 - ._ "’ : C-60 o L Wave Rotor Compression Ratio (PRw) Figure 43: Thermal efficiency, specific work, and SF C for the wave-rotor—topped engines versus the wave rotor pressure ratio and overall pressure ratio, Case A consideration 89 Overall Pressure Ratio Tu=1116.5 K npc=0.83 =0.82 =0.98 C-30 Relative Increase of Efficiency and Specific Work in “/5 Relative Change of Specific Fuel Consumption in “/0 Wave Rotor Compression Ratio (PRw) Overall Pressure Ratio 1',=1227.5 K $50.86 =0.82 =0.9e e96 C-60 Relative Increase of Efficiency and Specific Work in % Relative Change of Specific Fuel Consumption in % Wave Rotor Compression Ratio (PRw) Figure 44: Relative values of thermal efficiency, specific work, and SF C for the wave- rotor-topped engines versus the wave rotor pressure ratio and overall pressure ratio, Case A consideration 90 Case B: As described before, Case B considers another way to implement a wave rotor beneficially. While keeping the overall pressure ratio of the topped engine equal to that of the baseline engine, the compressor pressure ratio is reduced in the wave-rotor- enhanced engine. Lowering the compressor pressure ratio usually leads to a higher isentropic compressor efficiency (for comparable aerodynamic impeller quality, here simulated by using the same polytropic compressor efficiency), less the mass of the compressor, and probably lower manufacturing costs. Figure 45 depicts the actual T-s diagrams of the baseline engines and the wave-rotor- topped engines for Case B. Now, the overall pressure ratio is fixed at 3.6 and 4.8 for the C-30 and the C-60 engines, respectively. It is evident that the compressor work of the topped engine is less than that of the baseline engine. However, the turbine work is less too, but the heat addition for the topped engine is greater than that of the baseline engine. So, without calculating the thermal efficiency and specific work it is problematical to determine whether the topped engine has a higher performance than the baseline engine. Clearly shown in Figure 45 is that the turbine outlet temperature is considerably higher than that of the baseline engine (T 53> T 51,), making this case attractive for heat recovery applications due to the available additional thermal energy from the exhaust gas. 91 1500 1200 900 600 Temperature [K] 300 1500 1200 900 600 Temperature [K] 300 . 0"0'40'5b baseline engine 0-1 A-Z A-3 (4 {5.1 topped engine E Tllrbjne ________ h 0 G30 Engine 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [Id/kg K] . 04545-55 baseline engine 0-13-23-33-43-53 topped engine 3 e L 1.1-dine ___________ . _______ . _ p In 53 10 o C-60 Engine 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [kJ/kg K] Figure 45: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case B implementation 92 Values of the thermal efficiency, specific work, SFC and as well as their relative increase of the wave-rotor-topped C-30 engine can be obtained from the plots in Figure 46 and Figure 47, respectively. Similar to Case A, both thermal efficiency and specific work are monotonically increasing functions of the wave rotor pressure ratio PRW while SF C is decreased. Now, the relative increase of thermal efficiency (dash dot in Figure 47) is considerably less than the relative increase of the specific work (dashed), because the heat addition is greater in the wave-rotor-topped cycle. Similar results are obtained for the C-60 engine, not shown here. For Case B, the results show that with a wave rotor pressure ratio of 1.8 the thermal efficiency of the C-30 engine increases from 14.9% to 15.8%, and from 24.8% to 27.3% for the C-60 engine. Similarly, the specific work increases from 128 kJ/kg to 150 kJ/kg for the C-30 engine, and from 116 kJ/kg to 137 kJ/kg for the C-60 engine. Finally, SF C of the C-30 engine decreases from 0.156 kg/kN.s to 0.147 kg/kN.s and from 0.094 kg/kN.s to 0.085 kg/kN.s for the C-60 engine. For the same PRW = 1.8, wave-rotor-topping of the C-30 engine gives a relative increase of 5.9% and 17.1% for the thermal efficiency and specific work, respectively, and 5.6% reduction in SF C. For the C-60 engine, the performance improvement is less with 9.7% and 18.0% increase of thermal efficiency and specific work, respectively, and 8.7% reduction in SF C . 93 Compressor Pressure Ratio 3.6 3.0 2.57 2.25 2.0 1.8 T"=1116.5K 1; 0.28 :' 0.28 ”#:0‘83 2' : np,=0.82 i 0.24— 0.24 ” =9” en 2 A i » 3’ c 0.2 - 0.2 \ .9 ’ '1 a ’ a 5 g 0.16- 130.16 180 g 2 ~ ,3 3 O 0 o 0.12— m 0.12 160 5 a Z 8 if - o. o 0.03 _- 0.08 140 w E . 8 _ o. 0.04 - 0.04 120 "’ C C-30 0 '- 0 100 Wave Rotor Compression Ratio (PRw) Figure 46: Thermal efficiency, specific work, and SF C for the wave-rotor-topped C-30 engine versus the wave rotor pressure ratio and compressor pressure ratio, Case B consideration Compressor Pressure Ratio .\° 3.6 3.0 2.57 2.25 2.0 .s 0- C _ .9 .\° 3 ' ..\° .5 E -6- E 15 2 - 5 ;° ° 5 u ‘1’. 1 1,; s .2 '12" m a a - “6 2 .3 _ 3‘, o a. ” iii 0, ~18- e a “5 ~ E 8 0 k a; 5 2’ E o w -24— 5» 2 .c 0, i3 0 - [z 3 E - n: N a -30- 1! Wave Rotor Compression Ratio (PRW) Figure 47: Relative values of thermal efficiency, specific work, and SF C for the wave- rotor-topped C-3O engine versus the wave rotor pressure ratio and compressor pressure ratio, Case B consideration 94 0 Comparison of Case A with Case B The performance enhancement of Case A and Case B can be compared visually by using the plots in Figure 48 and Figure 49 for both C-30 and C-60 engines, respectively. For Case B the results are shown with dashed lines and the corresponding compressor ratio is shown in the upper scale in orange. The corresponding overall pressure ratio for Case A is shown in purple and the results are shown with solid lines. Case A clearly shows a more beneficial performance enhancement than Case B. 0 Cases C, D, and E The actual T-s diagrams of the baseline engines and the wave-rotor-topped engines for Cases C, D, and E are shown in Figure 50 to Figure 52, respectively. More detailed documentation of these cases are not presented here. The reader is referred to Ref. [230]. Instead, numerical values of the predicted performance enhancement of all five investigated cases with a wave rotor pressure ratio of 1.8 are summarized in Table 2. In this table, 177 represents the turbine pressure ratio. Subscript “gain” indicates the relative increase of thermal efficiency and specific work and decrease of SF C. Table 2 shows that Case A gives the highest performance increase for both baseline engines. Afier Case A, Case D gives the highest overall performance for the C-30 engine as for the C-60 engine. However, Case B provides the second highest thermal efficiency and the lowest SF C for the C-60 engine. Figure 53 and Figure 54 show maps of the relevant design space for Cases A, B, and D for each engine. The only fixed parameters are turbine inlet temperature, the polytropic efficiencies of the compressor and turbine corresponding to the respective baseline engine, and the combustion chamber pressure loss as indicated in the upper right corner legend of each map. Performance maps valid for Cases C and E of the C-30 and the C-60 95 engines are shown in Figure 55 and Figure 56 which have lower turbine inlet temperatures than that indicated in Figure 53 and Figure 54. Instead, Cases C and B have the same combustion end temperature as the baseline engine, as indicated in the upper right hand corner of these maps. 96 .___ Panm 3.60 3.00 2.57 2.25 2.00 1.80 Pale 3.60 4.32 5.04 5.76 4.68 7.20 : T“=1116.5K A 0.28 7 0.28 ”30.33 240 "3 1 =0.82 Z . i 0.24 — 0.24 ‘ 220 m _ “ 1 a E 0.2 — 0.2 200 f .2 ' fl ‘6. I 3 i‘, _ E g 0.16_ 2 0.16 180 g m . ,2 g g ' t u o 0.12— No.12 160 E a : g :1 _ D. '3 0.08- 0.08 140 w E I 0 - a 0 04 - 0 04 120 "’ : C-30 0 L 0 100 Wave Rotor Compression Ratio (PRW) Figure 48: Thermal efficiency, specific work, and SF C for the wave-rotor—topped C—30 engine versus the wave rotor pressure ratio and overall pressure ratio for Case A (solid) and versus wave rotor pressure and compressor pressure ratio for Case B (dashed) .__— PNIPl0 4.80 4.00 3.43 3.00 2.67 2.40 PaIPm 4.80 5.76 6.72 7.68 8.64 9.60 I Tu=1227.6K 0.28 — 0.28 ,1 =0“ 240 if : ”=0 32 5 ‘ =0 98 \ 0.24 _- 0.24 ' 220 g : Pl"? ’5 c 0.2 - 0.2 200 f .2 ' 'i a : a 5‘— g 0.16 r E 0.16 180 1; In ‘2 g g H t u o 0.12— LU0.12 160 E 3 - 3 if i a u 0.08- 0.08 140 «n 1.: . .5 C 3 0 04 f o 04 120 "’ i C-60 0 ‘- 0 100 Wave Rotor Compression Ratio (PRW) Figure 49: Thermal efficiency, specific work, and SF C for the wave-rotor-topped C-60 engine versus the wave rotor pressure ratio and overall pressure ratio for Case A (solid) and versus wave rotor pressure and compressor pressure ratio for Case B (dashed) 97 1500 . 0‘lb'45‘58 baseline engine o‘lc'zc'3c'4c'5c topped engine 1200 Z 3 900 h 5 ti 5: o. g 600 5.. 300 . 0 C-30 Engine 0 l I J l I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [lei/kg K] 1500 . 0‘18'45‘58 baseline engine 0--lC-2C-3C-4C-5C topped engine _ Tim!“ _________________________ 1200 ' i 2 900 ' :1 E". 8. g 600 - [— 300 ' . 0 C-60 Engine 0 I I J L I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [kJ/kg K] Figure 50: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case C implementation 98 1500 i 0‘10'48'55 baseline engine 0-10-20-30-49-5.) topped engine 1200 2 2 900 5 E 8. g 600 1- 300 . 0 C-30 Engine 0 I I I I I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [Id/kg K] 1500 . 0-1,,-4,,-5b baseline engine 0'10’20‘30'40'50 topped engine 3:) -Tivrlnne ____________________ 1200 r 2 2 900 " 3 3 0 o. ,E, 600 - [- 300 ' . 0 C-60 Engine 0 L I I I I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [id/kg K] Figure 51: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case D implementation 99 1500 . 0-1b-4b-5b baseline engine 071E72E73E74E75E topped engine 1200 :Tlnrbine . 3E 4b 7777777777777777.9477/3777 . . E I \ _ 1’ a z: 1 ,, 900 - ,’ '- \ I- I ._ 5 g ’z . 5 g V’ ” 5E :- 2 ” ‘5 600 - :1 ’.. 1- . ., a ’ Iii-lip“ I 300 - ’ . 0 C-30 Engine 0 J I I I I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [Id/kg K] 1500 r0'15'4b'50 baseline engine mix-2,73,34,75E topped engine _T1urbine_. _ _ __ , it 2: _ 1200 - ' '7 7 7"7'7'7 7 7 7 7 :211‘747; 7 7'7 9. E! \ If. \ E. 1’ i‘\ “ 900 ' ’1‘”, I '1‘. a .r” ’ \ a .1 ,I s, E 600 ' g ” [7' lb=lE 6’ I I .l I 300 ' . 0 C-60 Engine 0 I I I I I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [Id/kg K] Figure 52: T-s diagrams for the C-30 and C-60 wave-rotor-topped engines for Case E implementation 100 m... a..." 3. en” a..- w. 2 an ..3 an... E. its :43 Q3 3.: :.3 w.:. v.3 :.~ Wm 3.:N N.m~ TX; ...ax0km: 1am 13 - 0:1 ...mm 1.- 2:3-.- 19w. -. ..Sl ...mm -wdw..- ..fiES: v3: 33.: run: 33.: man: v3: «3.: 33.: :3: :3: «3.: ::~.: b . . :3: 03.: «3.: 03.: :3: 03.: 33.: mm—.: 33.: 53.: w«:.: 03.: TthxUkw v3 :3 :3 >3 VNN m3 n3 :3 ::~ :3 man 33 Rose: a 6m mam 1:3. mom . .3 m3 wowl :N «mm EN 2% :mm .mwxévz...“ 18..“ -- We... . 8m .. Wei 3.. ..m. of: - 12.1 ...e: .. 1.....1 18m 3:. : .33.... :28- mm.m .EM vmé ween mm.m :NM mmd :.v..m ::.N .36 wmé .Q - om... 3m om... 5m -Pa emu.-- SN. -8...“ . SN 18W 9.... . 3m E mm: mm: mm: Sun .3: van :3 mm: .20:— wv: .30 mm: Rose: 5 mam. 3: mm: 5:. man. 3: :v: 3:. mam. 3: mam. E: ETR . :90 :m-0 :90 :90 :90 :90 :90 :90 :90 :90 :90 :90 359.0 .95. v5 .QEB .95. 3.... 2.35 20:33:80 5 .2... 2.5.3. .2... 2.5.3. 3:023 258%. 2.5.3 o .ofinnEoQ 5 one. 3 .8353 .83.:an o .30.: :m.o>o o 5.0.3.:an o H 030 n 030 0 02.0 m 8.0 < 030 :. _ :0 or... 0.3.3.: .80. 02.3 a £3» m=_&o.-.o.o.-o>a>» mo momma oz). :5. moEwco 2......me 5952. camcaano oocmEStom ”N 033. 101 Rcomp=3.6 1:1 C-30 Engine Tu=1118.5 x 240 ' n,c=o.33 :5 Info-32 2' mangoes 220 X 5’ - E 200 g ‘u‘: . 3 X 180 g ' 180 a? ‘ 8 - 8 , ‘ 14o ' 120 - ' 1oo 1 2 3 4 5 6 7 8 9101112131415 Compressor Pressure Ratio Figure 53: Performance map for wave—rotor-topping of the C-30 engine, Cases A, B, and D Rcomp=4.8 C-60 Engine ‘ \7 _ ..., 7.512275 K _ 11 "$.36 '1. 11,155032 ~ [Imam ,. '—$: 4- " 220 .l. src (kg/kN.s) Specific Work (kJ/kg) No Gain Region 120 100 1 2 3 4 5 6 7 8 9101112131415 Compressor Pressure Ratio Figure 54: Performance map for wave-rotor—topping of the 060 engine, Cases A, B, and D 102 Rcomp=3.6 | C-30 En ine Tu=1116'5 K I g 11,5033 24° ‘5? No Gain Info-32 2' I Region “co-r5033 220 i‘ | g’ A 3 l 200 g “a“: 2 180 f O 3 0 E O 8 U) Compressor Pressure Ratio Figure 55: Performance map for wave-rotor—topping of the C-30 engine, Cases C and E Rcomp=4.8 C-60 En ine T =1227.6 K g n:c=o.85 240 src (kglkN.s) Specific Work (kJ/kg) 12 3 4 5 6 7 8 9101112131415 Compressor Pressure Ratio Figure 56: Performance map for wave—rotor-topping of the C—60 engine, Cases C and E 103 The maps allow predicting the performance of the wave-rotor-enhanced engine in terms of thermal efficiency (green), specific work (blue), and SF C (red) for any combination of the compressor pressure ratio (abscissa) and the wave rotor pressure ratio PRW (parameter labeled in black). In these maps, the multiplication of compressor pressure ratio p,,/po and wave rotor pressure ratio PRW determines the overall cycle pressure ratio [7,/pg (orange). The locus of optimum compressor pressure ratio points (for highest thermal efficiency and specific work at each achievable wave rotor pressure ratio) are connected by black solid lines. The optima for SF C are found at the same combination of the compressor pressure ratio and PRW as the optima of the thermal efficiency. Such maps are not only very useful to explore the possible enhancement of already existing baseline engines, but they also serve well for selecting a design point or region for designing a new wave-rotor-topped engine. In all plots, the performance points of the baseline engine (PRW=1) and the wave-rotor-enhanced engines of all cases with a wave rotor pressure ratio of PRW=1.8 can be found. For instance in Figure 53 and Figure 54, starting from the performance point of the baseline engine, the performance values for Case A are found by moving vertically upwards (e.g. along the dashed line for constant compressor pressure ratio p” /po) until the corresponding performance curve of the expected wave rotor pressure ratio is crossed. Case B is found by moving along a line of constant overall pressure ratio pg /po (orange). The results indicate that for every compressor pressure ratio in each design space shown here, the performance of the topped engine is always higher than that of the corresponding baseline engine with the same compressor pressure ratio (Case A 104 consideration). The increase of PRW always increases the performance. However, for higher compressor pressure ratios the benefit of using a wave rotor progressively diminishes. In fact, for compressor pressure ratios greater than around 11, almost no benefit can be obtained for the C-30 engine. An identical statement applies to the 060 engine for compressor pressure ratios above around 15. The benefit is clearly the greatest for lower compressor pressure ratios. This suggests that the wave-rotor-topping for microturbines with low compressor pressure ratios can produce the greatest relative benefit. Moreover, as expected and known for baseline engines (PRW =1), it is also true for wave-rotor—topped engines that the compressor pressure ratio for the maximum specific work is always less than that of the maximum thermal cycle efficiency. However, with increasing wave rotor pressure ratio, the optima come closer together, while moving towards lower compressor pressure ratio. This can be viewed as an additional advantage for applying wave rotors to small gas turbines with low compressor pressure ratios. So as the plots show, adding a wave rotor with a 1.8 pressure ratio to C- 30 or C-60 baseline engines with a compressor pressure ratio p,1/po= 3.6 or 1),, /pg= 4.8 respectively, already brings the design point into the optimum range for highest specific work and nearly half way closer to the optimum for highest thermal efficiency. 4.2.4 Comparison Between Adding a Second Compressor Stage with Wave-Rotor- Topping The wave-rotor-topping competes mainly against adding a second compressor stage to the single stage baseline engine. In this competition one major advantage of the wave- rotor—topping is that the wave rotor favorably operates mechanically independent from the high-speed engine shaft. Therefore, adding a retrofit wave rotor does not require the redesign of the challenging dynamic system. Even if the compressor or turbine wheel is 105 adapted subsequently to utilize the full potential of the wave-rotor-topping, the dynamic system may change but not as dramatically as if a second compressor stage was added. Thus, by default the wave rotor is a system for achieving similar thermodynamic advantages as by adding a second compressor stage or a high pressure spool, but with many fewer dynamic challenges. To justify the wave-rotor—topping approach further, the performance results of both competing solutions are compared below. For the addition of a second compressor stage, performance data are calculated for the five most probably relevant pressure ratios of the second stage described here: ' 1762 =PRW A (perhaps) logical way to compare both systems would be to assume the same compression ratio for the second compressor stage as for the wave rotor. Hence the compression ratio of the second compressor stage would be 170:1.8, because the assumed wave rotor compression ratio is PRW =1.8. ' "’0: WC] More likely, when the effort of adding a second compressor stage is undertaken, the designer would not limit the pressure ratio of the second stage to 17c2=l.8. It might be desired to a add second compressor wheel that is similar to the existing first stage (or the same) for reasons such as using existing experience, or producing both wheels cost effectively as identical wheels, or producing them geometrically similar using the same or slightly modified tools. This approach can be modeled by setting the compressor shaft work of the second stage equal to that of the first stage, simulating the same angular momentum change of the flow in both stages at the same shaft. Because the inlet air 106 temperature for the second stage is much higher, the pressure ratio of the second stage is less than that of the first stage 17C2<17c1 . ' 1762: 1761 Alternatively, it could simply be assumed that the pressure ratio of the second stage is equal to the pressure ratio in the first stage. This is a common design approach. ° 17c (w)opt. and ”C(71)”; It might be desirable to compare the wave-rotor-topped engine with a two-stage compressor engine that has an overall pressure ratio 1752 '170 corresponding to the optimum for maximum specific work 17c (w),,,, or the optimum for maximum efficiency 17c (Wopr- The resulting values for specific work and thermal efficiency respectively are the maximum values actually obtainable by enhancing the pressure ratio of a conventional compressor. The values of 17(w)op, and INmOP, can be found by using performance maps in Figure 53-Figure 56 and following the curves for PRW =1 to their optimum points. In all performance calculations above, it is assumed that the polytropic compression efficiency of the second stage is equal to that of the single stage baseline compressor. The performance values of all these two-stage-compressor cases as well as intermediate cases can also be read off the performance maps in Figure 53 to Figure 56 following the curves for PRW =1. The compressor pressure ratio at the abscissa then corresponds to the overall pressure ratio 170170. The performance results are compiled for the C-30 engine in Table 3 and for the C-60 engine in Table 4 for all five two-stage-compressor cases described above. The two-stage-compressor engines are compared with the wave-rotor- topped engine Cases A and B. These cases are more suitably compared with two-stage- 107 compressor engines for a few reasons. Both cases employ the same compressor as the first stage of the baseline engines. Case A has shown the highest performance improvement and it represents the maximum performance achievable for a wave-rotor- topping cycle. In Case E, the baseline compressor is the same and the combustion end temperature is the same as for the baseline engine, not requiring any thermal enhancement of the combustor. It is understood that this is exactly the case for the two- stage-compressor engine, where the combustion end temperature is simultaneously the turbine inlet temperature (which is never the case for wave-rotor-topping). This has been illustrated in Figure 57 that compares the T-s curves of the baseline cycle, the modified cycle with the two—stage compressor, and the wave-rotor-topped cycle Case E with a wave rotor pressure ratio of 1.8 for the C-30 engine. The two-stage-compressor values may also be compared with the wave-rotor-topping Cases B, C and D using the supplied data in Table 2. Table 3: Performance comparison between adding a conventional second compressor stage and wave-rotor-topping Case A and E with a wave rotor pressure ratio of 1.8; baseline engine C-30 C-30 Two-Stage Compressor Wave Rotor Topped Baseline Feature 17cm)... 17C3=PRw17c(rz).p.lvcz=wc, 170:1}, Case A Case E 170 1.52 V 1.8 2.22 2.42 3.6 1.8 1.8 - fluff/ESL ‘_ 13.6 135, , 129 .131 1.35 all]. .331--. ‘28 SFCIkg/st] 0.133 0.128 0.126 0.127 0.141 _ 0.116 0.126 0.156 77 0.175m “0.181 0.184 0.183 0.164 0.200 0.184 0.149 C (10),... [%] 6.3 . 5.5 0.8 -23 -25.8 33.8 7.7 (81:0,... [%] 14.5 17.9‘l18.9 18.6 9.6 25.2 18.9 (1])gain[%] 17.5 21.5 23.4 22.8 10.1 33.8 23.4 108 Table 4: Performance comparison between adding a conventional second compressor stage and wave-rotor-topping Case A and E with a wave rotor pressure ratio of 1.8; baseline engine C-6O C-60 Two-Stage Compressor Wave Rotor Toppedl Baseline Feature 17C(w)0,,,. f HC2=PRWHCQ])_OE_WC2=WC, ”C2=17Cl Case A Case E 170 1.52 1.8 2.55 2.79 4.8 1.8 1.8 w [kJ/kg] 193 192 179 173 118 231 190 184 SFCIkg/kN.s] 0.105 0.101 0.098 0.098 0.115 0.095 0.102 L 0.120 F.._--_fis._-1u._fi _ , - - ‘ _ ,.1-- fl-.___-1.2 a- A.“ 1_ 77 0.222 0.229 0.236 0.236 0.203 0.242 0.227 0.194 (w)ga,-,, [%] 4.9 4.3 -2.7 -6.0 -35.9 25.1 2.9 (SF C)ga,-,, [%] 12.5 15.8 18.3 18.3 4.2 20.1 14.9 (7] )gum [%] 14.4 18.0 21.6 21.6 4.6 25.1 17.5 1500 ' 0-1 1:41:51, baseline engine 0'12'23'35'42'51: topped engine 0'11'41'51 two-stage compressor engine 1200 :.TL'!!|1-1¢._.. ......... , ---,-.- _____ _ ________ x _. I E. ’ , a 900 - / ’_,/ :5 ’ I ’ I a ’ ’Jg'fi'! ’ ’ l5 1: ’ "J .J’ ’ I 2' 600 - 2 ‘5’”), I ’ a: E a a ’ 111:] 2:11 ’ 300 " 0 G30 Engine 0 I I I I I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [Id/kg Kl Figure 57: T-s diagrams for baseline cycle, conventional cycle with two-stage compressor (double compression work) and wave-rotor-topped cycle Case E with PRW =1.8, C-30 engine 109 Table 3 and Table 4 show that the gain in predicted thermal efficiency, specific shaft work, and SF C of the wave-rotor-topped engine in Case A is always greater than any obtainable values for the two-stage-compressor engine. A look at the maps in Figure 53 and Figure 54 clearly verifies this for the relevant design space. In these plots, the performance points for Case A lie well above any point at the curves for PRW =1 where all the performance data of the two-stage-compressor engine can be found. For the C-30 engine, Case E in Table 3 still shows a higher performance than any two-stage- compressor configuration. For the C-60 engine, however, Case E in Table 4 achieves nearly the same performance as a two-stage-compressor engine with a second-stage compression ratio in the range between the two optima for maximum specific work and maximum thermal efficiency (minimum SF C), 17c2=1.52. .255. Finally, the results show again that the compressor pressure ratio for the maximum specific work is always less than that of the maximum thermal cycle efficiency (minimum SF C). Besides the drawbacks of the two-stage-compressor implementation already mentioned, a second compressor stage adds not only a second compressor wheel, it always requires an enhanced combustor capable for higher combustion pressure and a turbine adapted to considerably a higher pressure ratio. Finally it requires an enhanced engine shaft, transmitting much more compression work. This situation is quite opposite for the wave-rotor-enhanced engine in which the transmitted compression work for all considered wave-rotor-topping cases is always either less (Cases B, C, and D) or the same as for the baseline engine (Cases A and D). The combustion pressure ratio can be kept the same (Cases B and C). Furthermore, the pressure ratio that has to be accommodated in the turbine is always less for the wave-rotor-topped engine than for the 110 two-stage-compressor engine with the same or greater overall pressure ratio, likely causing fewer problems when adapting the turbine. For the wave-rotor-topping Cases B and C, the turbine pressure ratio is even less than that of the single stage baseline engine. Additionally, in Cases C and E the wave-rotor-topping even lowers the turbine inlet temperature, which allows the designer to use a turbine made of a cheaper material. 4.2.5 Substituting the Compressor in the C-60 Engine with the C—30 Compressor plus Wave Rotor An interesting practical engineering option is to substitute the current compressor of the C-60 microturbine with the smaller and cheaper compressor designed for the C-30 microturbine by adding a wave rotor to obtain the same overall compression ratio of 4.8 required for the C-60 engine. This case is similar to the Case E shown in Table 2, only the compressor polytropic efficiency is switched from 85.9% for the C-60 compressor to 82.9% for the C-30 compressor. Also different from the Case E, the wave rotor pressure ratio needs to be adapted to 1.33 (instead of the value 1.8 in Table 2) to obtain the C-60 baseline overall pressure ratio of 4.8 in combination with the C-30 compressor that has a compression ratio of 3.6. For such a modified C-6O engine, Figure 58 shows the thermal efficiency, specific work, and SF C as a function of wave rotor compression ratio (lower abscissa) and compressor pressure ratio (upper abscissa) if the overall compression ratio is fixed at 4.8 and the compressor has a polytropic efficiency of 0pc =0.829 of the C-30’s compressor. It is seen that by using a wave rotor with a pressure ratio of 1.33 (orange points), the modified C-60 engine would have an thermal efficiency of 19.1%, a specific work of 190 kJ/kg, and SF C of 0.122 kg/kN.s. Compared to the C-60 baseline engine (yellow points where PRW =1) with a thermal efficiency, specific work, and SFC of 19.4% and 184 kJ/kg, and 0.120 kg/kN.s, respectively, the topped engine would not 111 enhance the efficiency and SF C, but the specific work would increase by about 3%. Furthermore, this engine would use a smaller compressor that has a lower manufacturing cost and already exists. Other, more advantageous combinations with an even smaller but newly designed compressor and higher wave rotor pressure ratio are also conceivable, as shown in Figure 58. Compressor Pressure Ratio 4.8 4.0 3.43 3.0 2.67 2.4 F T f r r r r r r r 0.28 l 0.28 ---,--Pa/ 9 3'5 “21:3“ 240 7»? . 1 "P" ' z : n,,=o.82 f 0.24 b 024 _.__.._...,..-_ .. ..... - ””5033 220 a ’— A i r 11 \N ’4 P" 1: 0.2 _~ 0 2 - _ _ . . (w ~ _ 200 a a: L. 5‘ —————:f",/ -1 3 I. L. I ""'”""".T_.: §O.16_ .§°16/ I nan-g 8 i E '7 ' src i E o 0.12 mm 0.1-2 m7 160 E - ” ‘ 0 § 2 | 1 a 0 0.08 - 0.08 ~ ~ 1 ~~~~~~~~ ._ 140 w E C | I 8 - . . 0.0 04 - 0.04 — ~- 120 01 ~ | . : palplfl=!1.33 : ° ' 01 1.2 1.4 1.6 1.8 2”" Wave Rotor Compression Ratio (P R“) Figure 5 8: Thermal efficiency, specific work, and specific fuel consumption of the wave- rotor-topped C-6O engine using the C-30 compressor versus wave rotor compression ratio and compressor pressure ratio for an overall pressure ratio fixed at 4.8 (yellow points for original C-60 baseline engine, orange points for wave-rotor-topped C-60 engine with C- 30 compressor) 4.2.6 Effect of Compressor Inlet Temperature It is well known that the performance of gas turbines is affected by varying the ambient conditions [231]. For instance, both the thermal efficiency and output power of a gas turbine engine with a fixed turbine inlet temperature decrease when the compressor inlet temperature rises. This is disadvantageous if a stationary gas turbine is installed in 112 hot-weather locations. Under these conditions, it is proved in the following that the wave- rotor-topping is even more advantageous. For simulation it is assumed that the compressor isentropic efficiency stays constant although a slight decrease might be expected. It is fiirthermore assumed that the specific compression work stays constant since the compressor geometry does not change. Hence, the compressor pressure ratio decreases upon increase of the ambient temperature from basic thermodynamics. For the baseline engine it is obvious that this results in a lower turbine pressure ratio and lower specific work produced by the engine. The results for compressor inlet temperatures of 250 K, 300 K, and 350 K are compiled in Table 5 for the C-30 engine and in Table 6 for the C-60 engine. The corresponding performance values of the wave-rotor—enhanced engines of Cases A and B, which are the most considered cases discussed in the literature, are also shown in these tables. For the wave-rotor-enhanced engines the general trend is the same. However, while the absolute performance degrades at higher ambient temperatures for baseline and topped engines, the performance gains of the enhanced engines relative to the baseline engines increase, making this technology more desirable for applications under hot- weather conditions. For a range of wave rotor compression ratios PRW =1...2, this reversed effect is visualized in Figure 59 and Figure 60, especially in the lower part showing the relative gains of the topped engines. While Figure 59 shows the absolute gain and the relative performance enhancement for Case A, Figure 60 illustrates similar results for Case B. 113 Table 5: Ambient temperature effect on performance - comparison between baseline engines and Cases A and B of wave-rotor—topping with PRW =1.8; baseline engine C-30 C-30 To: 250 K To: 300 K To: 350 K Cases Baseline CaseA CaseB Baseline Case A CaseB Baseline CaseA CaseB w [kJ/kg] 168 217 190 128 171 150 95 132 117 SFC[kg/kN.s] 0.125 0.097 0'112 0.156 0.116 0 147 0.195 0.141 0.179 g A- ,- .. f - ,, .1 2 _ - ,. - ,. w“ .. A- - 1 . m. 77 0.185 0.239 0'219 0.149 0.200 0.158 0.119 0.165 0.130 (w)gm-,, [%] 29.0 13.0 33.8 17.1 38.6 23.1 (SFC)g,,,-,, [%] 22.5 3.6 25.2 5.5 27.8 8.6 ( r] )gm-n [%] 29.0 3.8 33.8 5.9 38.6 9.4 Table 6: Ambient temperature effect on performance - comparison between baseline engines and Cases A and B of wave-rotor-topping with PRw =l.8; baseline engine C-60 C-60 To: 250 K To: 300 K To: 350 K Cases Baseline Case A Case B Baseline Case A Case B Baseline Case A Case B w [kl/kg] 234 285 258 184 231 208 144 185 168 SFC[kg/kN.s] 0.010 0.082 0'39 0.120 0'29 0.117 0.144 0.113 0 139 PM.-. 2. -2 2. -~_ a. 2*. ._ _7-- ,H2 2. A" #71 gMMum A—- -1. # A .1 fl ,7 0.232 0.283 0'723 0.194 l 0'224 0.199 0.161 0.206 0.167 (w)g,,,-,, [%] L 21.8 10.5 25 l 13 0 28.2 16.4 (SF C)g,,,-,, [%] 17 9 1.8 20.1 2.6 22.0 4.0 by _n_. 1 _ __ .fi ..7.. ..1.___._fi LA ..- .k- ‘_-._A (77 )8“... [%] 21.8 1.8 25.1 2.7 28.2 4.2 114 Specific Fuel Consumption ( kg / kN.e) r,=111s.5 K ‘ T_=1227.s K 0-23 ' nwsoes ' ' nwaoas =0.82 =o.az 0.24 - =o.ea . - =o.9e 0.2 - >~ 2 0.15 - a ‘2 e- on — w . 0.08 — 0.04 _- o h. Wave Rotor Compression Ratio (PRW) Wave Rotor Compression Ratio (PRW) 60 T“=1 116.5 K T“=1227.6 K 11,630.86 50 30.82 40 30 20 0 Relative Change of Specific Fuel Consumption in "I. Relative Increase of Efficiency and Specific Work in "/2. Wave Rotor Compression Ratio (PRw) Wave Rotor Compression Ratio (PRW) Figure 59: Effect of ambient temperature: absolute and relative changes of thermal efficiency, specific work, and SF C versus wave rotor pressure ratio for Case A 115 Specific Work (kJ/ kg) Specific Fuel Consumption ( kg l kN.s) 1 r =111s.5 K Tu=1221.6.5 K 0.28 — “ ‘ ’ ' qw=o.es . . =o.az 0.24 — . . =o.9a 0.2 r 0.16 L 0.12 — 0.08 - 0.04 _- 0 1. Wave Rotor Compression Ratio (PRw) Wave Rotor Compression Ratio (PRw) g c 0 — 0 E ru=111s.5 K ru=1227.s.5 K .g . . 71,5086 1: ..\° g -2 .s 25 =o.9a in _ >. C U 0 C 9 .4 _ g 20 , .- .2 f. g -6 — $15 0 u: a » a m 8 ‘6 -8 - 510 81 . ,3 C 5 g a 0-10 - g 5 G .2 * .. %-12 - 0 I! Wave Rotor Compression Ratio (PRw) Wave Rotor Compression Ratio (PRW) Figure 60: Effect of ambient temperature: absolute and relative changes of thermal efficiency, specific work, and SF C versus wave rotor pressure ratio for Case B 116 Specific Work (kJ I kg) Relative Increase of Specific Work In % 4.3 Gas Turbine with Recuperation Figure 61 shows a block diagram of a recuperated gas turbine topped with a four-port wave rotor. The wave rotor is placed after the compressor and before the recuperator. Figure 62 and Figure 63 show T-s diagrams for the C-30 baseline engine and the corresponding wave-rotor-topped engine for Cases A and B, respectively. Path O—lb-Zb'~ 4b—5b-5b' represents the baseline cycle and path 0-11-21-21.-31-41-51-51. (i=A and B) indicates the wave-rotor-topped cycles. Only Cases A and B are considered in this study due to their high potential for performance improvement under recuperated conditions. All calculations have been presented only for the C-30 engine, while the conclusions are valid for the C-60 engine, too. Combustion Chamber Wave Rotor Compressor Turbine Figure 61: Schematic of a recuperated gas turbine topped by a four-port wave rotor Compared to the baseline engine without the recuperator, the recuperated baseline engine here has a lower turbine inlet pressure due to the pressure loss of the recuperator air-side. However, it has a higher turbine outlet pressure due to the pressure loss of the recuperator gas—side. 117 Temperature [K] 1500 1200 900 600 300 F0-1b-2b'4b-5b-5b' baseline engine 0-1 A-ZA-2A33A-4A-SA-5A' topped engine: Case A b L Tartan _ -, _ b - 1A . o . C-30 Engine 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [kJ/kg K] Figure 62: T-s diagrams for recuperated baseline and wave-rotor-topped cycles, Case A Temperature [K] 1500 1200 900 600 300 '0-1b-2b'4b-5b-5b' baseline engine 0-13-23-23333-43-53-5,’ topped engine: Case B C-30 Engine l I l l I 1.5 1.8 2.1 2.4 2.7 3.0 Entropy [Id/kg K] Figure 63: T-s diagrams for recuperated baseline and wave-rotor-topped cycles, Case B 118 4.3.1 Thermodynamic Calculations In this section, calculations have been performed for both Cases A and B of the C-30 engine equipped with a recuperator. As before, in the wave-rotor-topping cycle, it is assumed that the compressor inlet condition, turbine inlet temperature, compressor and turbine polytropic efficiencies remain unchanged and are the same as for the baseline engines. For the recuperated cycle, the pressure losses across the air and gas sides of the recuperator are assumed the same and equal to 2% (”recap—air = 17mm... = 0.98). The effectiveness of the recuperator is considered to be ”recup=90%- These values have been selected based on typical existing recuperators. The major challenges in providing a recuperator with greater effectiveness are size and cost. For stationary applications, size and weight are not critical, but for mobile applications these limitations produce recuperator effectiveness commonly less than 90% [224]. Complying with Figure 61, the following steps are used to calculate the thermodynamic properties of the gases in different states of the topped cycle [232]: 0 Path 0-1: Compressor Equations (21) to (24) are again used to calculate the air properties at the compressor exit. 0 Path 1-2: Compression in the Wave Rotor As shown before, Eqs. (25) and (26) are used to calculate the air properties after the wave rotor compression process. 0 Path 2-2‘: Recuperator Air-Side The recuperator effectiveness based on the actual and maximum heat transfer from the turbine exhaust gas to the air can be expressed as: 119 T T, (2. ’5 T,—T, I I 77...... = (48) Using this equation, the temperature at the exit of the recuperator air-side (Tm) becomes: 713' 2 T!) + ”ramp (715 — 7.12) (49) In this equation, the turbine outlet temperature (7)5) is still an unknown parameter. More equations are required to find the unknown values Ty and T,5 . These additional equations will be derived later. The corresponding pressure at state 2‘ is: :17 pr2zn I p12 recup~uir recap—air Po p0 PRW 17.: (50) where I],e,.,,p_a,-, = pm /p,2 represents the pressure loss across the air side of the recuperator. 0 Path 2:3: Combustion Chamber Similar to Eq. (27), the following equation is used to calculate the firel/air ratio: ”thPR:(1+f)Cpgas 773—Cpair712' (5]) where T, 3 is an unknown value. However, f can be alternatively expressed as follows: 77g f hPR = (1+ f)Cpgas 7:4 '1' WWE _ WWC _ Cpuir (7:2 " 7:2) _ Cpair Tn (52) Since the net output work of the wave rotor is zero, f can be expressed as: _ Cpgus 774 -Cpair (712’ _772)_Cpair711 f (53) 77g hPR T Cpgas T14 This expression also will be later used to find 732*, T,5 , and f . Now, the total pressure after the combustion chamber is obtained by: b. : £1521.— : ”comb ”recap—air PRW 17C (54) p0 pt). p0 120 0 Path 3-4: Expansion in the Wave Rotor The wave-rotor characteristic equation (38) is used to calculate the turbine inlet total pressure expressed in Eq. (41). 0 Path 4-5: Turbine The turbine total outlet pressure is: EL: 1 p'5' = I (55) po 17 recoup-gas p0 17 recap—gas where Hrecupgafi p,5¢/p,5 represents the pressure loss across the gas side of the regenerator. It is justified to assume that the total pressure of the gas leaving the recuperator (p.54) is equal to the ambient pressure ([70). The turbine specific work according to pressure ratio across the turbine is calculated by using Eq. (42) where pg must be substituted by p,5. By obtaining the turbine specific work, the turbine total exit temperature (7)5 ) is given as: r... -l w p /p T —T - T =T — T I— —’5—-1 7“" 56 Now, to find values of Ty, T ,5 , and f , it is necessary to use Eqs. (49), (53), and (56), along with Eqs. (38), (41), and (55). The procedure is now explained. Substituting Eq. (49) into (53) gives: _ Cpgus 774 _ I7rer'up Cpair (715 _ Ti} ) — Cpair 77’ ”Q hPR _Cpgas 7‘14 f (57) or, 121 Cp gas 77 - ( h _ C as 77 )— C air 71 TU=TH+ _ 4 f 779 PR pg 4 p l (58) I7recup Cpuir Now, equating this equation with Eq. (56) results in: -I Cp...T. -f(77 h —Cp..T, )—Cp..,T, , / + 1, 4 Q PR 1: 4 1:714-777 T74 1_(P5 pa) 7,... (59) ”recap Cpuir pH / p0 Using Eqs. (41) and (55) gives: + Cpgus 774 - f(,]Q hPR _ Cpgas‘ 7.14 ) _ Cpair Til ”remap Cpair I ’_w_’ (60) =Tr4-777T14 1-( )rg... recap—gas POHC where PO is a function of f and is obtained from Eq. (3 8). The above equation computes f as a function of other known cycle parameters. 0 Path 56‘: Recuperator Gas-Side The temperature of the gas leaving the recuperator (T ,5.) can be calculated by using the energy equation across the regenerator as follows: Cpair (712’ —7;2):(1+f)Cpgas(]-;5—T;5°) (61) This yields: C . T . —T T . = 7:5 _ pair ,2 12 (62) ’5 Cpgw 1+f Now, it is possible to calculate the engine performance parameters as described before. 4.3.2 Predicted Performance Results Implementing Case A for the C-30 recuperated engine, Figure 64 illustrates the variations of the cycle thermal efficiency (dash dot), specific work (dashed), and SF C (solid) with increasing wave rotor pressure ratio PRW. For the baseline engines (without 122 the wave rotor, PRw = 1), whereas the thermal efficiency of the recuperated cycle is much greater than that of the simple cycle (without the recuperator) its specific work is slightly less due to the pressure losses across the recuperator. With a conceivable wave rotor pressure ratio of 1.8, the thermal efficiency of the recuperated engine increases slightly from 24.8% to 26.6%. However, the specific work increases from 116 kJ/kg to 160 kJ/kg. The thermal efficiency of the recuperated cycle remains almost unchanged when the wave rotor is used, because the temperature difference between the air and the gas entering the recuperator of the topped engine (Tm-Tm) is much less than that of the baseline engine (Tab-m). This results in reduced recuperation, and hence the greater heat addition into the topped engine offsets the increased work output. This is in contrast to the unrecuperated engine where both the baseline and topped cycles have the same heat addition, as explained before. SF C for the recuperated engine slightly decreases from 0.094 kg/kN.s to 0.087 kg/kN.s. Again the increased heat addition into the recuperated topped cycle keeps SF C almost constant compared with the simple cycle, which is seen is Eq. (47). The relative increases of thermal efficiency, specific work, and relative decreases of SF C are shown in Figure 65. Even higher than the unrecuperated engine discussed before, the specific work of the recuperated cycle shows a significant improvement of about 37.6%, but thermal efficiency and SF C are only about 7% greater than those for the baseline engine. 123 Overall Pressure Ratio _ 3 6 4.32 5.04 5.76 6.48 7 2 e I T =1116.5K l I 1 i V V V I 1 V l 3 0.36t 0.36 ":50“ :190 - 2 1] =0.82 — E 0.32 r 0 32 flim=0.98 . _: 180 30.28} 028— - - 11 -—‘170 3 " * ...,..- - ----- ‘ o ’; f 5 0.24 ~ 0.24 :- rrrrr ~ - e, 0 ’ - 160 a ‘15. C 3 o’ i 1‘5 C 1’ _ X g 0.2: '8 02— «ml—,4 ~ 1150 3 c I E v j 3 8 0.16:- Lu016— .7. 1140 [g 8012: 012—.1’7’1-97 ‘130E I; ' * r SFC - a? g 0.08;- 0.08 ’4 -_ 120 O ' 2 D. r _ m 0.04 - 0.04 — ~ - 11o : C-30 wrth Riecuperator : '_ 1 1 1 1 1 . 1 1 1 l 1 1 1 1 1 L‘ o o, 1.2 1.4 1.6 1.8 210° Wave Rotor Compression Ratio (PRW) Figure 64: Thermal efficiency, specific work, and SF C of the enhanced C—3O recuperated engine versus the wave rotor pressure ratio and overall pressure ratio for Case A Overall Pressure Ratio Relative Increase of Efficiency in °/o Relative Increase of Specific Work in % Relative Change of Specific Fuel Consumption in % Wave Rotor Compression Ratio (PRw) Figure 65: Relative values of thermal efficiency, specific work, and SF C of the enhanced C-30 recuperated engine versus the wave rotor pressure ratio and overall pressure ratio for Case B 124 Similar to Figure 64 and Figure 65, now Figure 66 and Figure 67 show the absolute and relative increases of the thermal efficiency (dash dot) and specific work (dashed), and the decrease of SF C (solid) upon implementing Case B to the C-30 recuperated engine. Similar to Case A, both the thermal efficiency and specific work are increasing functions of PRW while SF C decreases. For PRW = 1.8, the thermal efficiency of the simple cycle is increased from 24.8% to 27.3%. Similarly, the specific work is increased from 116 kJ/kg to 137 lekg. Finally, the SF C is decreased from 0.094 kg/kN.s to 0.085 kg/kN.s for the recuperated cycle. Therefore, wave-rotor-topping of the recuperated engine gives a relative increase of 9.7% and 18% for the thermal efficiency and specific work, respectively, and an 8.7% reduction in SF C . 125 Compressor Pressure Ratio 3.6 3.0 2.57 2.25 2.0 1.8 : T =1116.5K 0.36 — 036 u 190 ,u? : 11,5083 5 0.32 _— 0.32 180 E. i ...... i‘. 0.28: 0.28 170 :25 E: 0.24:- 0.24 160 :, 3 02?~§°2 Hog 50m; Eme 1403 U . : LU . 5% e » 8 : 012- 012 130 g ‘1“. i "’ 1% 0.08; 0.08 120 w r- 6} 0.04 3 0.04 110 o L o 100 Wave Rotor Compression Ratio (PRW) Figure 66: Thermal efficiency, specific work, and SF C of the enhanced C-30 recuperated engine versus the wave rotor pressure ratio and compressor pressure ratio for Case B Compressor Pressure Ratio .\° 3.6 3.0 2.57 2.25 2.0 .5 ° 7 0 E - 2 s 2'1 ~ s .E S -6- -E e 2 3‘ g 8 - E u a 2 % _ - it: u:- 12 E (E- 0 ' o E u- 8 - 3 g -18 — e 3 s - E E 0 a: E 8’ E c» 19 ~24 - s E .1: q, I! U Q: T) .‘2’ - 1: N 3 -30 - 1:: Wave Rotor Compression Ratio (PRW) Figure 67: Relative values of thermal efficiency, specific work, and SF C of the C-30 recuperated engine topped with a wave rotor versus the wave rotor pressure ratio and compressor pressure ratio for Case B 126 Table 7 summarizes the results showing a comparison between performance improvements for these for implementations of both Cases A and B. It is seen from the table that the baseline engine with recuperation has a much higher efficiency (about twice) and lower SF C than those of the unrecuperated engine due to the heat addition reduction in the combustion process. Implementation of Case A into the baseline engine results in a significant performance improvement of the unrecuperated engine, however, the thermal efficiency and SF C improvements of the recuperated engine are lower than that of the specific work. In Case B, the topped cycle without the recuperator still benefits from the wave rotor even though the performance enhancement is less than that of Case A. However, the topped recuperated cycle has a higher performance compared to the unrecuperated engine and its thermal efficiency and SF C gains are even more than those of Case A. The results clearly demonstrate the benefit of implementing Case A for unrecuperated engines and implementing Case B for recuperated engines. Among all topping cycles, the best relative performance improvement is obtained with Case A topping of the unrecuperated engine. A comparison between the recuperated baseline engine and the topped simple cycle engine (PRW =1.8) reveals that for the specific example, the recuperator enhances the thermal efficiency and SF C, while the wave-rotor mostly enhances the specific work output. Therefore, substituting a recuperator with a wave rotor may be guided by a preference for high power output and reduced unit cost, considering that a recuperator contributes about 25-30% to the unit cost and a wave rotor may be cheaper. Finally, topping a recuperated gas turbine with a wave rotor can increase the performance especially if the topping cycle operates with the same turbine inlet temperature and same 127 overall pressure of the baseline engine (Case B implementation), which is preferable for the combustor and fuel injection design. Table 7: Performance comparison between baseline engines and two cases of wave-rotor- topping with a wave rotor pressure ratio of 1.8 17L, at *LclkL/kgi W .552 #99337 0.266 37,-6 . 6:2, , 7.2 (SFQML . 4.4 Turbojet Engines Topped with Wave Rotors Figure 68 illustrates how a four-port wave rotor is used to top a simple turbojet engine (without afterburner). Point “a” here refers to the ambient condition at the inlet diffiiser. Thermodynamic analysis of such a topped cycle is different form that of a stationary topped cycle due to the presence of the inlet (diffuser) and nozzle and the fact that in a turbojet cycle the net output work is zero, whereas in a grounded gas turbine the turbine work is greater than the compressor shaft work. 128 Combustion Chamber Diffuser 1| 0 \ l/ Compressor Turbine Wave Rotor Figure 68: Schematic of a turboj et engine topped by a four-port wave rotor Figure 69 visualizes all five wave rotor implementations in a schematic T-s diagrams. Path a-O-lb-4b-5b—6b represents the baseline cycle and path a-O-li-Zi-3i-4i-5i-6i (i=A, B, C, D, E) indicates the wave-rotor-topped cycles. Temperature a--0-1b-4b~5b-6b baseline engine a-0-1i-2i-3i-4i-5i-6i topped engine i= A, B, ..., E Tturbine Entropy Figure 69 : Schematic T-s diagrams for a baseline turbojet cycle and five different wave- rotor-topped cycles 129 4.4.1 Thermodynamic Calculations For the thermodynamic calculations, a reasonable case considered here is an aircraft equipped with a simple turbojet engine flying at an altitude of 10,000 m at Mach 0.8. The component performance parameters of the C-30 engine are used as the baseline engine. Other component efficiencies are listed in Table 8. Now, a higher pressure drop of 5% in the combustion chamber (17mm = 0.95) is considered in the calculations. It simulates a higher pressure drop in the combustor due to a possible size reduction of the combustion chamber which is attractive for small propulsion systems. At the altitude of 10,000 m, ambient air enters the inlet diffuser at Ta=223 K and [20:26.5 kPa . Table 8: Turbojet baseline engine data, assuming T 0 =300 K, Cpair=l .005 kJ/kgK, Cpgas =1.148 kJ/kgK, yair=l.4 , yga, =1.33 Baseline Engine Turbine inlet temperature T“ 11165K Compressor pressure ratio pub/3,0 3.6 Diffuser isentropic efficiency 770 93% Compressor isentropic 0 efficiency ”C 80 /° Compressor polytropic o efficiency ”PC 83 /° Combustion efficiency ”Q 98% Combustor pressure loss Hm”, 0.95 Mechanical transmission 0 efficiency '7” 99 /° Turbine isentropic efficiency m 84% Turbine polytropic efficiency mar 83% Nozzle isentropic efficiency my 95% 0 Path a-0: Inlet Diffuser With the given flight Mach number M, the stagnation temperature across the diffuser (Tm=T,0) is: 130 =Ta[I+ZL”—_—1M2) 2 (63) Using the definition of the diffuser isentropic efficiency (:70), the diffuser outlet total pressure is obtained from the equation : 2 7a” / fair '1 p10 u - = 1+ ——“——— 64 p“ [ 770 2Cp T] ( ) air a - Path 0-1: Compressor The approach used before is implemented for this stage. 0 Path 1-2: Compression in the Wave Rotor The approach used before is implemented for this stage. 0 Path 2-3: Combustion Chamber The approach used before is implemented for this stage. 0 Path 3-4: Expansion in the Wave Rotor The approach used before is implemented for this stage. 0 Path 4-5: Turbine In turbojet engines, considering the mechanical transmission efficiency, the compressor shaft work equals the turbine output work: Cpair(Til—7~10)=17M(1+f)cpgas(7~14—T'15) (65) Therefore, the total temperature of the gas leaving the turbine can be calculated as: C . T -T, 715:7,” pair( (I 0) (66) — m. (I + f) 6p... To find the total pressure of the gas leaving the turbine (p,5), the value of the turbine isentropic efficiency (In) is needed. There are two ways to calculate 771 for a turbine: 13] (&)(7gas‘l)’ll'r”Iran" _] p, 77f : I: . ( f5 )(rgm-l)/7gac _1 p14 and, T5 _r_ -1 (TM) 777 : (fii)7ga§_’/7gut _1 pm It is preferred to use the turbine polytropic efficiency mar to obtain p,5 equating the above two equations, one finds: rgm [9,5 =(Ir_5_)(7,....-I)npr T p14 :4 or, ygm ££:!fi1(£ pa 1).. T (4 )(rgm —I)’II’T 0 Path 5-6: Nozzle For a given local pressure ratio at the nozzle exit (pig/pa), it is true that: 19.5 = p... 12.5 g = 10,4 19,5 /p.. 1 pr 12.. p” p. p. I’M/p. p. /p.. (67) (68) . Therefore, by (69) (70) (71) Parameter p,5/p6 is a useful quality for calculating the local gas temperature leaving the nozzle (T6) by using the definition of the nozzle isentropic efficiency (771v): 73m" ." gas p15 /p6 — d T6 : T15 _77NTr5 I32 (72) Since T,5=T,6 , the nozzle exit velocity (u6) can be calculated from the expression: “6 : JZCpgas (7:5 — T6) (73) Therefore, the nozzle exit Mach number is obtained: u M, = 6 (74) vygas Rgas T6 Finally, for the given pé/pu , the ratio of p,6/pa is calculated from the expression: y _ 1 7m“ f1:_p’_6_p_i=(1+ 5““ M62)7w'"1p_6 (75) p. p. p. 2 p. After calculating the thermodynamic properties of all states in the cycle, it is possible to evaluate the engine performance parameters. For instance, the thrust produced by the engine is given by: T=(mai,+m/)u6-marrua+(pa-Pa)A.v (76) where A N is the exhaust area of the nozzle. Thus, the specific thrust (thrust per unit mass flow of air) can be written as: Av ST=(1+f)u6-ua+(p6—pa)m‘ (77) air The pressure term in the above equation can be presented [233] as: RmT/T 1— / )g 6 a .0. 106] (78) R' u6/aa yair ("r A (p.-p.) .” =a.[(1+f m where an is the sound velocity of the entering air. Therefore, the specific thrust can be calculated as: \R805 7:5 /Ta 1—pa/p6 (79) IR' u6 /aa rair 0” ST=(I+f)u6 —ua+au [(144 133 The thrust pressure term (pé-pa)AN is non-zero only if the exhaust jet is supersonic. In this study, all calculations have been performed for pé/pu =1 resulting in zero value for the thrust pressure term. The specific fuel consumption is given by: SFC = i- (80) ST Finally, the overall efficiency is derived as: n - ST ”“ (81) 0 m. 4.4.2 Predicted Performance Results A detailed documentation of the results is not presented here. The reader is referred instead to Ref. [234]. However, numerical values of the relevant cycle data and the performance enhancement are summarized in Table 9, where indicates an attractive relative performance improvement in overall efficiency and specific thrust of about 15.4% and a 13.3% reduction in specific fuel consumption using a wave rotor pressure ratio of 1.8. Table 9 shows that for Case A perhaps the best combination of high overall efficiency, specific thrust and low specific fuel consumption is achieved. However, Case B provides the lowest value of the specific fuel consumption and the highest overall efficiency. 134 Table 9: Performance comparison between baseline turbojet engine and five cases of wave—rotor-topping with a wave rotor pressure ratio of 1.8 Case A Case B Case C Case D Case E o compressor ° 3:3” 0 turbine inlet 0 compressor Same as _ ratio I ocombustor pressure. Baseline baseline - turbine . turbine 0 turbine inlet - combustor engine inlet temp. 7 inlet IEIDP- temp. 7end temp. @[K] 7 1116.5 1116.5 7 1056.2 1116.5 7 1042.75 7 717(7 7 7 3607 7 72.070 7 7 290 7 2.65 7 7 73.670 7 7 177T 7 7 1.73 7 717.307 71.32 1.47 7 1781 wc [kl/kg] 140 68 68 100 140 §UNikgl B .562 z, Eli z 7493' 550,, a.-- $131 §fglgg0<101kPa , T3=1199.3 K , u2=111.8 m/s ,p,r= 4.45X101kPa , and T]'=510.8 K. The results can be further used to calculate the gas penetration length (nearly exactly L,,,.). It is an important design parameter which will be used to determine the required rotor port widths. The maximum possible value of the compressed air mass flow is reached when the gas penetration length becomes equal to the rotor channel (L,,., /L= l). The two following equalities are used to find the dimensionless penetration length L,.,,/L: L w —E/i+ "“5 (87) u, V u2+a3 W L L—L , 7_3+ gas gas 77:-*7 W (88) Equation (87) results by equating the time required for the air/gas interface originating from the lower comer of the inlet port to be overtaken by the expansion wave (LN/uz) with the opening time of the inlet port (W3 /V) plus the time required for the expansion wave to catch the air/gas interface (L,,../u2+a3). Tuning conditions TC 1 and TC 3 result in Eq. (88). Note that when the expansion wave meets the air/gas interface, its velocity reduces from u2+ a3 to u2+ 02. Multiplying both sides of Eqs. (87) and (88) by er and dividing them by length L and simplifying gives: 144 M5 M+a’ Lgus 0] 89 L — 1 W, ( ) W, 1 1 + _ M(1+M,) M+$ a; M+ 0.» a, a], a], where M: 112/01' and M3 = u; /a3 are the flow Mach number and the Mach number of the burned gas in region 3, respectively. M5: w5 /a,r is the shock Mach number where ws represents the shock wave velocity. The only unknown term here is the port width ratio W 2 /W3 . It is obtained by writing the continuity equation across the combustion chamber which gives: m, = m, + ”'2! (90) or, p321, A, =p2 u, A2+n'1f (91) Because u2= m and A3 = W3 . z , A2: W2. 2 where z is the height of the channel, the above equation can be simplified to: £2.52 = 1 + f (92) [’2 W2 Here, it is assumed that m, = mm, = #2,. Other cycles can exist in which the mass flow rates "'2, and n’z, are not the same, but they are not considered here. By using the ideal gas equation, Eq. (92) transforms to: —l W7 - 7 R’05 .. : Ralr I; I = A £(I'l‘f) (93) W3 Rgus T3 (1+f) Rair T2 145 Thus, substituting this equation into Eq. (89) estimates the gas penetration length. L,..,./L as a function of the initial cycle thermodynamics and the port boundary conditions. For Case A and corresponding thermodynamic parameters listed in Table 10, a very low value of Lm/L =0.05 is obtained. Similar to Keller’s results [44], the present result indicates that for typical wave rotor pressure ratios, e.g. PRW < 3, the gas penetration length Lga,./L is very small, e. g. less than 0.2 for such tuned wave rotors. Hence a relative small amount of air compared to the wave rotor channel volume is delivered to the outlet port. Furthermore, it leads to very small and unrealistic port width W3 and W2 . For instance, inlet gas port width can be found by using Eq. (87) as: W3=Lg,, —1—-+ I V (94) uz u2+a3 Also, Eq. (93) gives the port width ratio Wz/W 3 from which the absolute value of W; can be obtained, by using W; from Eq. (94). In this study, W3 = 0.45 cm and W2= 0.21 cm are obtained for a rotor with a channel length of L = 20 cm and a rotational speed of V=50 rer. A huge (and unrealistic) channel height of H...” = 35 cm would be then necessary for the desired volume flow, resulting in an undesirably bulky wave rotor. An alternative solution should be sought to increase the gas penetration length. This is the topic of the next section. 5.1.2 Charging Process with Two Shock Waves Several wave diagrams can be introduced that employ two shock waves. Two of these wave patterns were introduced in Chapter 3 and they are studied in more detail in this section. As it will be shown in the following, considering a secondary wave inside the channels results in reasonable geometry parameters with a greater relative gas penetration length. 146 0 Charging Process with Two Primary Shock Waves Figure 73 depicts a wave diagram which employs two shock waves. It is similar to the previous wave diagram sketched in Figure 72, but both inlet and outlet port widths have been proportionally extended and a shorter rotor height is used, matching the same mass flow rate through the rotor as considered before. Also, the gas inlet port becomes much greater than the air outlet port keeping the same port width ratio as before to allow more gas to penetrate the rotor channels. " 'I Figure 73: Wave diagram of high-pressure part with two primary shock waves By closing the outlet port, a second primary shock wave originates at the upper outlet edge and propagates from the right to the left with velocity w5.a;, into the part of the channel that contains air flowing with velocity u; from left to right. It reduces the velocity of the air to zero to satisfy the zero-slip boundary condition at the end wall. After 147 crossing the air/gas interface, the velocity of the second primary shock wave increases from W50), to Magus, where the latter is the velocity of the shock wave in gas region 3. The increase in the velocity is mainly due to the fact that the burned gas has a higher temperature than the compressed air. The double-compressed air region and the compressed gas region after the second primary shock wave are introduced as regions 2' and 3', respectively. Here, a favorite case is considered where the closure of the gas inlet port is timed with the arrival of the second primary shock wave. This is called the tuning condition 2 (TC 2) [6, 40, 79]. Hence, expansion waves are not generated after closing the gas inlet port. To find the useful design parameters for such a wave diagram, all previous derived equations are used again, but new expressions must be derived for W3 , W2 , and L,” /L. Hence, the two following equalities are used to find the new value of dimensionless penetration length L... /L: “w = — + + g” (95) u 2 “’5 V WS-arr L L E1 7 gas + gas (96) V ”2 WS-gas The first equation is found by equating the time required for the air/gas interface originating from the lower corner of the inlet port to reach the second primary shock wave (Lga, /u2) with the time required for the first primary shock wave to run through the channel (L/ws) plus the outlet port opening time (W 2/ V) plus the time required for the second primary shock wave to meet the air/gas interface ((L-Lgas ) /w5-a,-, ). The second equation is obtained by equating the inlet port opening time (W3 / V) with the time required for the air/gas interface originating from the lower comer of the inlet port to 148 reach the second primary shock wave (Lga, /u2) plus the time required for the second primary shock wave to travel through the gas region and to meet the upper corner of the inlet port (Lgas/wsgm). Multiplying both sides of Eqs. (95) and (96) by all and dividing them by length L and simplifying gives: 1 4, L803 _ MS WS—gjr L 1 a, W, 1 a, —— +———— —-—— —- + M W3- W; M wS air —gas (97) In this expression, WSW and ws-gas are unknown and must be found. To find WSW-r , the strength of the second primary shock wave propagating in the air region 2 (175.0”: pzl/pz) is first obtained by using Eq. (7): 2 yair a / + I ”2 = —2_(US—air ‘" I) 7 _1 (98) yet? \ ”S-air + 2,0;— yair + 1 Then Ws.a,'r is found by using Eq. (9): yair + 1 WS-air = 02/ J—(HS-air —1)+1 (99) 2701? Finally, the corresponding Mach number can be calculated by using Eq. (1 1): . I MS—air :\/72mr + —(HS—air ‘1)+1 (100) yair A similar approach is used to find 175,33“,= p 3 ’/p 3 (strength of the second primary shock wave propagating in the gas region 3), WSW, , and M51305 using the gas properties. Therefore, Eq. (97) calculates the dimensionless penetration length for the wave diagram shown in Figure 73. For the case considered in this study L,.../L = 0.66 is calculated which 149 .‘\ is a considerably greater value than that calculated for the wave diagram with the single shock shown in Figure 72. The higher value of this parameter is due to the extension of the gas inlet port width, allowing more burned gas to enter the rotor channels. By obtaining L,,../L it is now possible to estimate the rotor geometry, as described before. By obtaining W 2 and W 3 using Eqs. (95) and (96), respectively, the inlet opening time (t3) and outlet opening time ((2) become: W3 ’3 :7 (101) and, t —— W3 (102) " V Additionally, the time required for the primary shock wave to run through the channel (II) is simply equal to L /w5. To find the other geometry characteristics, the following procedure is described. For a given baseline air volume flow rate, Q2 , the cross-sections of the outlet port 2 and inlet port 3 can be obtained from: A: = - (103) and, 23:21 A, = (104) where Q; is the volume flow rate of the burned gas, which is different from the air volume flow rate Q; . Considering the mass added from the fuel, the continuity equation relates Q2 and Q3 as follows: 150 p3Q2(1+f) (105) p3 Q3: The cross-section of the outlet port A2 is typically 10% of the frontal area of the rotor [6]. Therefore, the diameter of the rotor represented by 0,0,”, can be found from: 2 A2 E19. 75.91"“:— (106) 100 4 Assuming a rectangular cross-sectional channel, the channel height is given by: Han = ‘ (107) which is identical to H(.,.u=A3/W3 due to the continuity equation used in the derivations above. The rotational speed of the rotor (nmm), which is a variable to be chosen by the designer, is related to the rotor angular speed (co) by: n = — (108) where a) = V/Dmm/ 2 . Finally, the channel width can be found from the expression: ” D romr Wu” = -—N—— (109) where N is the number of channels for a one-row rotor, e. g., in the order of 30. Implementing the procedure explained in this study for Case A, the discussed channel and port parameters were calculated and are listed in Table 11 and Table 12. Results clearly show that the scenario employing two primary shock waves provides realistic results. Figure 74 shows the location and size of inlet and outlet ports of the charging process based on the results obtained in the present study. 151 Table 11: Flow properties inside rotor channels for L=20 cm Flow properties Two shocks Channel pressure in region 1' (p,) 4.45 XIOI kPa Channel temperature in region 1' (T2) 511 K Inlet and outlet flow velocity (:13 = 112) 112 m/s Outlet air Mach number (M2) 0.24 Inlet gas Mach number (M3) 0.17 Static pressure ratio across the primary shock ([75) 1.40 Velocity of the first primary shock (ws) 525 m/s Mach number of the first primary shock (Ms) 1.16 Static pressure ratio across the second primary shock in the air region (H5.a_,-,) 1.38 Static pressure ratio across the second primary shock in the gas refinflgw) 1.24 Velocity of the second primary shock in the air region (w5.,,_,~,) 436 m/s Velocity of the second primary shock in the gas region (wag) 630 m/s Mach number of the second primary shock in the air region (M5_,,_,~,) 1.15 Mach number of the second primary shock in the gas region (Mew) 1.10 Penetration length (L823) 13 cm Table 12: Wave rotor design parameters for L=20 cm, N=30, nmmr=10000 rpm, and m, =0.318 kg/s, Geometry and port characteristics Two shock Rotor tangential speed (V) 50 m/s Rotor diameter (Dung) 9.63 cm Cell width (Ween) 1.01 cm Cell height (Hail) 2.23 cm Circumferential length from 0 to opening of air outlet port (W 2) 1.92 cm Width of high pressure air outlet port (W2) 3.26 cm Width of high pressure gas inlet port (W 3) 7.02 cm Port width ratio (W 2/ W 3) 0.46 Angle from 0° until opening high pressure air outlet (61,) 22.85 ° Opening angle of high pressure air outlet (a2) 38.76 ° Opening angle of high pressure gas inlet (a3) 83.52 ° Time from 0 until opening of high pressure air outlet (’1) 0.38 ms Outlet air opening time (t2) 0.65 ms Inlet gas opening time (t3) 1.39 ms 152 (11:22.8 ° ‘7 t1 =0.38 ms 073 :835 o W1=1.92 Cm t3 =1.39 ms (12 =38,8 ' 7 W3 =7.02 cm 12 =o,55 ms '- W, =3.26 cm Gas Inlet Port Air Outlet Port Figure 74: Preliminary port designs of the charging process, corresponding to the wave diagram sketched in Figure 73 0 Charging Process with Two Shock Waves and an Expansion Wave Figure 75 shows a second different wave diagram that also utilizes a secondary shock wave, but now as a reflected shock wave. Compared with the wave diagram shown in Figure 72, this wave diagram has a shorter (more compact) wave rotor design as seen in the figure and a relatively great gas penetration length. Creating such a reflected shock wave always implies that the static pressure in the high-pressure air outlet port is greater than that of the high-pressure gas inlet port. This can easily be the case when using a combustion chamber with a higher pressure loss as discussed further below in Figure 77. Here, for the same combustor pressure loss as considered before, employing a reflected shock wave results in a shorter wave rotor and a reduced velocity of the leaving high- pressure air. Because of mass continuity, the high-pressure air outlet port is now wider. Figure 75 shows that with a reflected shock wave and a shorter rotor, the expansion wave arrives afier the upper outlet edge, if the tuning condition for the shock wave at lower outlet edge (TC 1) is kept and the inlet width remains unchanged. This also has been mathematically proved by Keller [44]. In this figure W2 is still the outlet port width 153 before shortening the rotor (in the absence of a reflected shock wave as discussed before). Here, W0", is defined as the outlet port width and WT is introduced as the outlet port width at which the tuning condition for the expansion wave at the upper outlet edge is fulfilled (TC 3). Thus, the actual outlet port width Wm), for the shortened rotor is wider. In fact, it is in the domain of W2 g W0", 3 WT. For a perfectly tuned design, Wm,r should equal W7, however it is highly possible that for a given PRW, TC 3 might not be achieved. Then, another shock wave would originate from the upper comer of the outlet port in order to satisfy the zero-slip boundary condition at the end wall. This is undesirable and not shown in Figure 75, although it describes a general case where TC 3 is not fulfilled. For such an untuned situation, the restriction W2 5 WW, 5 W T must still be satisfied. Figure 75: Outlet port width comparison before and alter rotor shortening In the shown case, to reduce the outlet velocity, the reflected shock wave originates at the lower outlet edge and propagates fi'om the right to the lefi. The double-compressed air 154 after the reflected shock wave leaving the wave rotor toward the combustion chamber is region 2. In Figure 75, L, represents the location where the reflected shock wave intersects the air/ gas interface. The compressed gas region after the reflected shock wave is named region 3'. The reflected shock wave reduces the speed of the flow leaving the channel from u2r to W and causes a further local pressure rise within the channel from p2' to p2. By fulfilling TC 2, the entire fluid column in the channel consisting of two different fluids moves to the right with the pressure and velocity of the outlet air. In this way, the entire fluid column is compressed by the reflected shock allowing more gas to enter the channel. Similar to the same approach described for the charging process without a secondary shock wave, again Eqs. (82)-(86) are used but with some modifications: Zyair ,/ . R. T, —._T “27 7 air rm , £377] 7,", +1 (110) ya” pl” _p_3+&'_’;1 pl/ rair +1 u ’ —u - ir T,,=T,+(2 “)2 (111) 2Cpair 70" ...,—l ££=[_]i_2]7 (112) [’2 T2 2 _T + ”2’ (113) I3 3 ZCPgas 7pm m-I &=[_Tg_]’ (114) P3 T3 where T2 is now equal to: 155 73:77:; .. [T 7.2 l T. T3 (115) T 2./ T ,4 and T2 /T2 , and functions of the local pressure ratio across the primary (175: p 3/p 1 .) and reflected (HIM-r: p2/p3) shock waves, respectively, and can be obtained by using Eq. (5). Furthermore, um), is the induced velocity caused by the reflected shock wave and can be calculated from Eq. (7) using 172-02,. Thus, 6 unknown parameters p 3, T 3 , u2' , pp, Ty, and 17mm exist for the above five Eqs. (110)—(114). This shows that introducing the reflected shock gives an additional degree of freedom. This allows selecting arbitrary values of 175 and [722-02, in a confined range. Thus, several solutions can be obtained for given engine data. However, if it is assumed that pp: p, and T1,: T 2 the number of unknowns is reduced and unique solutions can be obtained for 4 unknown variables p3, T 3 , u 2, , and 1722.“... This has been shown in a simple five-step procedure for a triple tuned wave rotor examined in our previous work [235, 236]. For example, Figure 76 shows the variation of 175 and 1722...), versus PRW for a given 170mb . It is clearly seen that, for the given [760”, higher PRW values can be only achieved by simultaneously increasing 175 and decreasing 1712-02,. By assuming pp: p, and T1,: T2 , the results also show that for a PRW greater than around 1.8, 172-0), becomes less than unity which is physically impossible for a shock wave. However, a higher pressure drop allows for a higher 1722-02, as it is shown in Figure 77 described below. 156 1.5 mm: 0.83 T“= 1116.5 K 1.8 1.4 1.6 / 1.3 1.4/ 115 1“IR 11m,,&"‘0';’9r 1-2 1.2 / 1.1 11.5 l 1.6 1.7 1.8 I 1.9 l 21 Wave Rotor Compression Ratio (PRw) Figure 76: Variation of 175 and HR versus PR W for a constant value of 1760,"), Assuming pp: p; and T 2': T, , Figure 77 provides a general map relating the shock strengths 175 and 1722-02, to various combinations of PR W and 110mb. The right boundary of the map is the 1712-02, =1 line, where the reflected shock wave vanishes. The lower and lefi boundary is the u2=0 line, simulating the closed wall condition where no flow leaves the channel anymore. This condition occurs when 1723.“), approaches the value of 175 . Beside zero exit velocity line, several lines of finite exit velocity are shown between the lefi and right boundaries. They show that for a given PRW, increasing 17mm causes a reduction in u2 due to the induction of mass motion behind the reflected shock wave. 157 A‘ 4mm." 3'0 (13.-51111111; I-i 71-497041-3 1-2 1.1 HR=I . .— " ‘ ‘15.? 200 7 l a I. " ‘ 0.03 I ‘ 1 275 _ I 100 T 7 \ - . - ‘ r ~ . [0.02 I , T ~ . 1.9 3 1.7\ . M 2.0 * ~ :4 opt' - - - - '. ' As/R=0.01 l ‘ ~ I g * ' - . - . .2 1.5‘ 1.5 — ' ~ . ._ ‘ 0.0025 ‘~ ~ ” "1:0 HS=L1 1.0 L 1 41 L l 1 J_ 0.6 0.7 0.8 0.9 1 mental) Figure 77: Shock strengths 175 , 17R , entropy production As/R , and exit velocity u2 versus P R W and [lamb The most interesting feature in the map may be the lines of constant entropy production in the channels due to the shock waves. The entropy production As is calculated by summing the entropy generations across the primary and reflected shock waves. They have been normalized by gas constants of the air and gas, respectively. For a desired fixed PRw , coming from the HM.) =1 line with only a primary shock, As/R reduces as the reflected shock 17R...” becomes stronger. This shows the potential for higher efficiency, if the wave rotor is designed with a reflected shock wave. However, this continues only until the optimum line (solid black) is reached. The optimum line indicates the combinations of PRw and [lamb at which the least entropy production occurs in the shock wave compression for a given PRw. Left of it, the entropy production 158 increases again for a given PRW. Simultaneously in Figure 77 from the right to the left, [760mb decreases. This allows for a combustor with a higher pressure drop that may be smaller and more compact, which may be especially desirable for aerospace applications. In fact, Figure 77 shows that for the desired PRW=1.8, the value of [Zomb=0.98 considered in the thermodynamic analysis gives a value of 17R-a),<1. Thus, a higher pressure drop (meb<0.98) should be selected for higher values of 1722-02,, which may preferably reduce the size of combustion chamber. Finally, the optimum line can be achieved at lower exit velocities, here lower than 100 m/s. After calculating the above unknowns, it is now possible to obtain an equation for the ratio of the outlet port to the inlet port width (Wow /W3) using the continuity equation at the inlet and outlet ports: W 1££L_L=1+f (no '02qu To find the gas penetration length for the charging process with the reflected shock wave, the following equation is introduced by equating the time required for the air/gas interface to be overtaken by the expansion wave with the opening time of the inlet port plus the time required for the expansion wave to catch the air/gas interface: 21:, Lg... 1; .L._«‘_Ls_ V u2+a3. u}. u, (up where 03' is the speed of sound in the hot gas region 3' after the reflected shock wave has passed through it and L, is the location where the air/gas interface meets the reflected shock wave. The temperature in the region 3' is found by using Eq. (5). The local pressure ratio across the reflected shock wave propagating in region 3 (1722-80,: p y/p 3) is calculated 159 1.7- similar to the method described for the charging process with the secondary shock wave. The value of L, can be found by equating the time required for the air/gas interface originating from the lower comer of the inlet port to reach the reflected shock wave (Ls/212') with the time required for the primary shock wave to run through the channel (L /wS) plus the time for the reflected shock wave to meet the air/gas interface ((L'Ls) /WR-air )3 L, L L - L5 - = — + (118) “2" W5 WRoair Therefore, LS/L can be obtained through this equation as: I l M +w /a L arr ,- L 15 R 1 ’ (119) —+ M WR_W / a 1’ Multiplying both sides of Eq. (117) with a], and dividing it by the length L leads to: AF" 1 ]_ W: L L M u,/a,, MRL gas: I L 1 1 (120) (u, + a3.)/a,, - u_, /a,,. where MR is the rotor Mach number defied as MR: Way. The value of W 3 / MR /L can be obtained from TC 2: W3 L + L—L‘ + L‘ (121) MR MS wkw, /a,, wk“ /al, which leads to: W" = I + I + I — I L‘ (122) MRL Ms wR_m.r/al,. w /a, wR_a,,/al, L R—gas l 160 Therefore, Lg“. /L can be obtained based on Eq. (120) for a given PRW and 172.0,“. Figure 78 shows the variation of Lg“, /L versus PRW for different values of 172mb. It clearly shows that the newly obtained value of Lem/L is much greater than that calculated before. Therefore, employing a shock wave increases the gas penetration length significantly. 1 nwc=o.03 ’// TM=1116.5K 7‘3/ ‘ 08 / ’A/Q’ISA” , , v.0, //‘ x09 a" _ /,/ "A “’6’? 06 1 -/ ," ’1’?“ ' ' ‘ . 0 i // ’I’d'I’ 00‘“ a ’ " .r" —J a’ ,e’ 0.4 a ,/ 0.2 Q1.5 1.6 I 1.7 I 1.8 l 1.9 I :2 Wave Rotor Compression Ratio (PRW) Figure 78: Variation of L,gm /L versus PRW and 17comb Finally, the maximum outlet port width WT for which TC 3 is satisfied can be obtained by equating the time required for the primary shock wave to run through the channel (L /ws) plus the opening time of the outlet port (W;- /V) with the opening time of the inlet port (W3 / V) and the time required for the expansion wave to reach to the end of thechannel: L L—L i+fl=fl+—gm_+ gas (123) ws V V u7,+a, u2+a2 161 By multiplying both sides of this equation by a), and dividing it by length L and simplifying: 1 Lg“, 1 1 I _ + ._ .— M5 L (u2+a2)/a, (u2+a3,)/al, (u2+a2)/a I W, Wj/MRL I" (124) As mentioned above, for any given PRW, W 7- Wow 20 should be a restriction. However, a designer may prefer to obtain a value of W2— Way, as small as possible to fulfill TC 3. Figure 79 shows the variation of dimensionless values of (WT— WWJ/ WT versus PRW for various values of n.0,... 100 um: 0.83 9° Tu=1116.5K so °\° 70 l.- ; 60 ”g 50 3 - 40 .- 5 30 20 10 qI.5 1.6 1.7 1.8 1.9 2 Wave Rotor Compression Ratio (PRw) Figure 79: Variation of dimensionless values of (W2- W0.,,)/ WT versus PRW and 17mm The plot shows that the expansion wave reaches the right channel end at different times for different combinations of PRW and [Zomb- Thus, for a desired fixed PRW there is only a unique 17mm), that satisfies TC3 and vise versa. Furthermore, with increasing 162 PRW the divergence from TC 3 increases, which is not preferred for an efficient design. In a practical design, the rotor may incorporate various cavities (pockets) in the end plates to reduce the sensitivity to a mismatch of the various wave arrival times [44]. These mismatches can be caused by other events such as variation in the engine load. 5.1.3 Validation To validate the accuracy of the analytical prediction model described above, comparisons with numerical results of a test case have been made. Very little experimental data are published in the literature concerning successful operation of wave rotors. Therefore, it was decided to validate the above analytical model with existing numerical predictions which have been validated with some previous wave rotor investigations. The data of the wave rotor numerically designed at University of Tokyo are selected [215]. Table 13 shows the pressure and temperature of the inlet and outlet ports of this wave rotor as they were imported into a computer. code written in MathCAD software according to the above analytical approach. Additional design parameters of the Tokyo. wave rotor are shown in Figure 80 in which HP stands for high pressure and LP is used for low pressure. Table 13: Inlet and outlet port data of University of Tokyo wave rotor, taken from Ref. [215] Inlet air pressure p,, 3.0 XI01 kPa Outlet air pressure p,» 10.1 XI01 kPa Inlet gas pressure pd 9.2 XI01 kPa Outlet gas pressure 712,4 3.7 XIOI kPa Inlet total air temperature 7",, 440 K Outlet total air temperature T,2 907 K Inlet total gas temperature 733 12048 K Outlet gas total temperature TM 973 K 163 7, 7.77 \\\\ 7 ,7’;// \“\‘7\ 7 \ 1 :1 I I’ \\ o " 1 " ‘1 — ----- 4——/ “° ----- 1 ex W l 7 1’ . ‘ /' \ 77 I 15920° - r \\ s 1‘ ’ ‘ .’ ° / 27' Gas-HP \\ I,” ii \ « \21-5 Air—LP X72: 7:. / n43° 128‘” .%77 / A'I'HP 11563” Gaflpisf 616 Charge Side (Left Side) Discharge Side (Right Side) Figure 80: University of Tokyo wave rotor, taken from Ref. [215] Using some of the data provided in Table 13 (cell number, cell length, mean rotor radius, and rotational speed) and port locations shown in Figure 80, the analytical model predicts reasonable wave rotor geometry parameters as listed in Table 14 in comparison with the published data of the Tokyo wave rotor The listed data of the model are obtained for initial channel pressure and temperature of p,» = 4.34 XI01 kPa and T]: = 707 K. Good agreement between the predicted results and those obtained by numerical 164 simulations is observed. The existing errors can be related to the assumptions made in the model, e.g., assuming pure air at the beginning of the high-pressure part of the cycle and using a constant specific heat at different states of air and gas. In the first step, the accuracy of the present model can be improved by solving the flow field in the low- pressure cycle to determine how much the temperature and pressure of the air filling the entire channel at the beginning of the high-pressure part model should be changed. Table 14: Comparison between the analytical model and the numerical data of the Tokyo wave rotor a ‘V Geometry and port characteristics Model Tokyo WR Cell width (If/egg) 0.49 cm 0.4 cm Cell height (H2221) 0.3 cm 0.3 cm Circumferential length from 0 to opening of air outlet port (W 1) 0.76 cm 0.88 cm Width of high pressure air outlet port (Wow) 1.65 cm 1.62 cm Width of high pressure gas inlet port (W3) 2.1 cm 1.87 Port width ratio Lng/ W 3) 0.78 0.87 Angle from 0° until opening high pressure air outlet (61,) 18.7 ° 21.6 ° Opening angle of high pressure air outlet (a2) 40.6 ° 40.0 ° Opening angle of high pressure gas inlet (a3) 51.7 ° 46.0 ° 5.2 Numerical Simulation Even though the above analytical procedure can provide some useful design parameters for a preliminary design, for a complete design, a whole-cycle analysis including the low-pressure part of wave rotor operation is required. In analytical approaches, the method of characteristics is often used to solve the flow field especially in the scavenging part of the cycle [98-105]. An alternative technique is to use CFD methods, allowing more accurate simulation of the flow field. It has been common practice to create and use specialized codes. As reviewed in Chapter 2, several precise numerical codes have been developed so far. However, most of these codes are not commercially available, hence a few groups of researchers who have developed these in- 165 house codes are using them, e.g., the NASA wave rotor code. A broader accessibility of appropriate CFD tools could facilitate a wide range of wave rotor analysis. The development of modern multi-purpose commercial software packages has reached a level that allows the successfiil modeling and analysis of the operation of many technical devices, including wave rotors. Several commercial codes are now available that can be applied to investigate many problems related to wave rotor design and operation. Such codes are particularly interesting for 2D and 3D modeling of full devices and some special problems. They offer tools that allow for relatively easy geometric preparation, a range of typical boundary conditions, relatively fast and robust solving, and a wide range of post-processing which is valuable for engineers and scientists. Yet, the use of such codes is not as fast as the use of common office sofiware. Although the geometry and boundary conditions can be modeled relatively fast, using such complex commercial software packages, the computational effort is still enormous, so that flow field is often only available after lengthy, time—consuming computations. Therefore, such simulations are not suitable for an initial geometry search or a geometry optimization but can be performed as a last stage of investigation, verifying solutions of particular problems or the full operation of a complete wave rotor. For preliminary investigations, initial design, and optimization they are not necessarily as efficient as specialized codes. Some attempts at simulating pressure wave superchargers with the help of commercial codes already have been undertaken. One such code is GT-POWER, in which pipe elements have been divided into a series of objects for which the conservation equations have been solved. An interesting description of techniques for optimization of 166 . -..- '_.. .....-. n timing, shaping, and control of pressure wave changers using GT-POWER has been described by Podhorelsky et al. [237]. The present work shows an attempt at developing 2D models of wave rotors using the commonly available CFD software FLUENT. This enables the realization of detailed gasdynamic phenomena inside the rotor channels. For most of the results investigated here, very little experimental data exist for possible verification. Therefore, the results are compared with some previous numerical data which are available for conventional wave rotors. The presented results shall not be interpreted as a proposition of a particular geometry for practical applications. While they can be a base for further investigations, they rather present some illustrations of physical phenomena generated in different configurations and phases of operation and showing the capability of the code. 5.2.1 Approach The numerically designed Tokyo wave rotor (see Table 13 and Figure 80) was selected as a model and reference for the numerical simulation presented here. The main reason for selection of this wave rotor is the availability of detailed geometry data and operating conditions. Figure 81 shows the temperature contour in this wave rotor predicted at the University of Tokyo. While for this port arrangement the pressure contour is not reported, it is available for another operating condition as shown in Figure 82. The goal here is to produce similar results using the commercial software FLUENT. The numerical procedure described here can be performed for almost all other wave rotor configurations. 167 "m 11".?“ mfiizkummww ,. “1111 kiiu W‘WIWM ’ . 1111'“ 1313.111113- m. ann1mlllllm1llllwllm “1m "‘an 7rmlItmumiuuunhlllii rotation Figure 81: Non-dimensional total temperature contour of the Tokyo wave rotor taken from Ref. [215] 168 IPrimr h kal Secondafl Shock Wave Air-LP rotation Figure 82: N on-dimensional total pressure contour of the Tokyo wave rotor, taken from Ref. [215] The first step in the flow simulation with FLUENT is the mesh generation for a particular geometry. GAMBIT is the default grid generator for FLUENT. Even though GAMBIT has the ability to draw the geometry, a more advanced drawing software like AutoCAD was used to sketch the geometry. Using AutoCAD significantly reduces the sketching work. After creating the geometry in AutoCAD, the geometry was imported into GAMBIT for the grid generation. The mesh size in the channels and inlet and outlet 169 ports was carefully selected to provide accurate and fast solutions. Tests showed that finer meshes would not provide more accurate results, only causing additional computational time. Figure 83 shows the unstructured mesh used for this study at the beginning of the wave rotor operation. In the computational model, 30 channels (not all shown) move in a block upward perpendicular to the ports. A finer mesh has been used in the interface of the channels and all four ports, as shown in Figure 84. This enhances the simulation of gradual port opening and closing. The most challenging step in generating the mesh is the creation of a stationary interface between the moving channels and the stationary ports. This was carried out by creating a relatively small gap (0.1 mm) between the wall ends of the channels and the end plates with the ports where an interface connects them. The yellowish line in Figure 85 indicates the interface where the entire right side connected grid lines slip along it during the wave rotor operation. An additional advantage of generating such a gap is the actual simulation of the leakage between the stationary end plates and the rotating channels. Once the grid was generated in GAMBIT, the rotor geometry was imported into FLUENT. Boundary conditions, the solving method, the working fluids, the convergence criterion, and several other functions (parameters) must be defined prior to the simulation. Air as an ideal gas was chosen as the working fluid. Such an assumption reduces the accuracy of the results because the effects of gas at the inlet and outlet ports and the burned gas in the rotor channels are neglected. However, the trends of the gasdynamic processes occurring in the channels are consistent with more accurate models as shown later. 170 ‘I Jul 19. 2004 Grid FLUENT 6.1 (2d. dp. segregated. lam] Figure 83: Generated mesh 1 1 Grid Jul 19, 2004 FLUENT 6.1 (2d. dp. segregated, lam] Figure 84: Finer mesh at port interface and channels, inlet high-pressure port 171 Interface / Grid Jul 19. 2004 FLUENT 6.1 (2d. dp. segregated. lam] Figure 85: Interface and gap between stationary ports and moving channels In this study, the governing equations of 2D unsteady laminar flow were selected and solved explicitly with a first order upwind difference scheme. The boundary conditions at the inlet and outlet ports play important roles in the flow filed simulations. These boundary conditions consist of flow pressure and temperature values at the inlet and outlet ports. The compressor exit pressure and temperature were assigned as initial values for the upward moving channels. 5.2.2 Numerical Results Figure 86 shows contours of total pressure at two succeeding time steps when the first moving channel is exposed to the high—pressure gas inlet port. As expected, the high- pressure flow starts compressing the low-pressure fluid in the channel by generating compression waves. The pressure stratification along the channel length indicates the action of compression waves. These compression waves form a stronger single shock 172 wave running through the channel. The bottom part of Figure 87 can be interpreted in this way. Figure 87 also shows how the second channel is influenced by the high-pressure gas inlet port as the computational time increases. Figure 88 depicts the moment when the head of the compression waves nearly meets the right end wall. The top figure indicates the moment before the wall contact, and the bottom figure presents the contact moment. The pressure contours clearly indicate the compression of the air by the compression waves. Once the compression waves reaches the end wall, a reflected shock wave is generated as shown in the two following time steps in Figure 89. The reflected shock wave compresses the air further. The doubled-compressed air leaves the channel by the opening of the high-pressure air outlet port. A pressure peak well above the inlet pressure is seen at the moment the high-pressure air outlet port opens. Figure 90 shows how the flow is scavenged from the rotor channels by opening the gas outlet port. This process is supported by the generation of expansion fans traveling through the channels toward the air inlet port. Finally, the fresh air is ingested into the rotor channels by opening the air inlet port as shown in Figure 91. It is interesting to note that a region of lowest pressure even less than the air inlet pressure (here shown numerically as negative pressure) is created just before opening the air inlet port. This can be interpreted as the reflection of the expansion wave off the left end plate shortly before the air inlet port opens. This low- pressure region significantly assists the ingestion process. Closing the air inlet port terminates the first operating cycle. The pressure contour at this moment is shown in Figure 92. All of the above described phenomena can be seen in this figure. 173 W. " ”at" ’fiaaiiififfinmelmw Figure 93 shows the enhancement of flow circulation in the beginning parts of the channels, which becomes dominant resulting in some outflow from of the channel shortly before the high-pressure gas inlet port closes. In the lower corner of the gas inlet port, the effect of gradual channel opening is magnified as shown in the top section of Figure 94, also showing some leakage effects. The bottom part of Figure 94 represents a focused zone of the upper comer of the high-pressure gas inlet port, indicating less leakage because of flow blockage caused by intense flow recirculation in the beginning of the channels. This flow blockage suggests that a narrower gas inlet port may be substituted for the current wider port. Figure 95, Figure 96, and Figure 97 show by colors the axial component of relative velocity vectors at air outlet port, gas outlet port, and air inlet port, respectively. For all three ports, recirculated flow regions exist in the upper port corner which again suggests that the port widths for all three ports can be reduced. Whereas the air inlet port shows the strongest recirculation, the air outlet port shows the least. For the above described figures only the first operating cycle is simulated. To reach a periodic solution, the code needs to be run with more cycles. For instance, convergence to a periodic unsteady state has been achieved after 20 rotations in a study by Fatsis and Ribaud [198, 201]. This can be performed in FLUENT in a 2D simulation by extending the current port arrangements, as shown in Figure 98 for a 3 cycle port arrangement (coarser mesh is used for a better visualization of the reader). Another approach would be a quasi 2D simulation with a 3D model Unfortunately, the generated data requires a huge memory, which was not available for this research. Such a periodic solution, however, 174 can be simply obtained in the 2D simulation of a radial-flow wave rotor. This is the topic of the next chapter. 175 lDle'US 9.75eou5 9.3913‘05 9.02e*05 , seams 3.84e‘05 3.47e‘05 3.10e'05 2.73e’05 Contours of Total Pressure (pascal) (Time= l. Bflflfle- 06] l08. 2004 FLUEN T6.l (2d. dp. coupled imp. lam.U unsteady) 1.026’06 9.8413415 9.48e‘05 9.12e‘05 8.7813415 8.40eo05 "Ti 8.05e'05 7.69e‘05 7.33e'05 8.976'05 ; saunas __, 6.256’05 5.90e*05 5.546‘05 5.18e+05 4.826'05 4.46e‘05 4.106‘05 3.75erfl5 3.3991‘05 3.036'05 Contoms of Total Pressure (pascal) (Time= 6. SDUUe-l] (1.8 2004 FL EN T8. 1 (2d. ldp. coupled imp. lam.U unsteady) Figure 86: Total pressure contours, the effect of high—pressure gas inlet port on the lst channel 176 l.l4e~06 l.lI]e*1115 ‘ 1.056416 1.016‘06 9.7Ue~05 9.256’05 8.86e'05 8.44e*|]5 8.036’05 7.619%15 7196415 750:1"; 6.776’115 ' 6.361345 5.941e‘05 1;: 5.526’05 5.111e*05 4.68e‘05 4.278‘05 3.85:.”05 3.43et05 3.1116'05 Contours of Total Pressure (pascal) (Time= 15111106 Jul [18. 211114 FLU ETN 6.1 (2d5. dp. coupled imp, lam. unsteady) 1.156’06 1.11er06 1.08e*06 1.0213416 9.78e005 9.32.3415 8.88e*05 8.43e*05 7.9913415 7.55e+[15 7.119'05 3.125415 2.686415 Centaurs of Total Pressure (pascall (Time=2.8110[1e Jul 118. 211114 FLU NT61(205.d]dp. coupled imp. larn. unsteady] Figure 87: Total pressure contours, the effect of high-pressure gas inlet port on the lst and 2nd channels 177 1,1 1.3-416 1.017e416 1.[|3e*06 9.878415 9.46er05 9.0E1er05 8.65e4)5 8.245415 7.846415 7436*05 7.03ev05 6.626415 6.21e‘05 5.816415 5.40er05 ; 4.99erfl5 4.59e‘05 4.18e‘05 3.77:415 3.37e415 2.96e'05 1111‘- Contours of Total Pressure (pascal) (Time= 3 3000e-0 5) Jul 08. 2004 FL LUENT 6.1 (2d. dp. coupled imp. lam. unsteady) 1.056'06 1.026'06 9.7923415 9.416‘05 9.03e'05 8658*05 8.27e415 7.89e*05 7.51e*05 7.1363905 6.7513415 6.37e415 5.9913415 5.611905 5.246415 4.86e415 4.48e415 4.106415 3.72e~05 3.34et05 2.96e*05 2%: Contours of Total Pressure (pascal) (Time= 3 9000e-0 Jul 08. 2004 LUNEN16.1 (2d. dp. coupled 1mp. lam. unsteady) Figure 88: Total pressure contours, traveling of compression waves toward end wall 178 7.655415 7.195.115 ': 61312.05 5; 6.2712415 ‘2 5.3151115 5.35.9415 “ 4.8913415 4.43e415 3.9713105 3.51e‘05 3.0413415 2.5812415 Contours of Total Pressure (pascal) (T1Fme= 4. 9000e- 05 1 Jul 08. 2004 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) 1.18er06 1.1 le*06 1.0713416 1.0312416 .1 9.84e'05 ' sums 8.9813415 85515.05 8.12e‘05 reams 7.2319415 6.8315415 6.40e+05 5.97e*05 5.54e+05 5.1115415 4.6812415 4.25e+05 3.82e405 , 3.3812415 " "i=r 2.95e+05 11 1 Contours of Total Pressure (pascal) (Time=5.70006-051 Jul 08. 21104 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) Figure 89: Total pressure contours, generation of reflected shock wave at two different time steps 179 .28e‘06 .23e416 .166416 .136436 .06e’06 1.036'08 9.84er05 9.35e415 8.85e‘05 6.366r05 7.87e*05 7.37e*05 6.86e415 6.39e‘05 5899*05 5.40e415 4.90e‘05 4.4le*05 3.92e*05 3.42e415 2.93e415 ._._-._.._..,4 an": (Time= 2. 0300e-0 Jul 08. 2004 Contours of Total Pressme (pascal) FTLUEN 6.1 (2d. ldp. coupled imp. lam. unsteady) 1.34e*06 1.298416 l.24e*06 1.19e‘06 1.13e*06 l.08e+06 1.03e'06 9.76e‘05 9.24e‘05 8.7113905 8.19e*05 7.666+05 7.14e+05 6.6]e‘05 6,096+05 5.576415 5.04e415 4.52e415 3.99e'05 3.47e'05 2.9449415 Contours of Total Pressure (pascal) (Time= 2. 2900e-0 41 U] 08. 21104 LUEN NT 6 1 (2d. dp. coupled 1mp lam. unsteady) Figure 90: Total pressure contours, air scavenging process at two different time steps 180 1.449415 1.36e'06 1.2913‘06 1.21906 1.146415 9 1.07e+06 5‘2: 9.93905 9.19e‘05 — 8.4551115 7.7169115 6.978905 8.236415 5.491905 4.756415 4.003905 ‘ 3.26e905 2.52e‘05 l.78c*05 1.04e'05 3.02904 -4.38e*04 Contours of Total Pressure (pascal) (Time-2790060412400 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) 1.346415 1.26e'06 1.1813006 l.lle*06 1.0313006 9.53e+05 8.77e'05 8.00e'05 7.23e*05 6.473905 5.70905 4.93905 4.16905 3.40e+05 2.833905 1.86er05 1.09e005 3.27e+04 -4.40e'04 -1.21e~05 'l.97e*05 3.5100 08. 2004 NT 6.1 6(204d )dp. coupled imp. lam. unsteady) Contours of Total Pressure (pascal) (T1me= Figure 91: Total pressure contours, ingestion of fresh air at two different time steps 181 L38e+06 l.31e+06 1.236416 1 1.15e+06 1.08e+06 1.01e+06 g _ 1 1 9.3le+05 1________ 1' f" f t; ' ' 7.80e+05 7.04e+05 6.29e415 5.5319415 4.78e‘05 . , 4.02e+05 7" " 3.2se+05 ; 2.51e+05 } 1.75e+053 ~ 7% 9 1"" r : 9.975414 1 -- .- .. " " l 2.41e+04 ' -5.15e+04 4.2712415 Contours of Total Pressure (pascal) (Time=3.9400e-041 Jul 08. 2004 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) Figure 92: Total pressure contour, end of the first operating cycle 6.23e+02 5.66e412 5.09e+02 ‘-: 4.51e*02 E 7.1 3.945412 .1 3.37e+02 2.79e412 2.22e412 l.65e+02 l.089+02 5.03e+01 ‘ 4‘056’00 Fit-5-" -:77. . . - 6.4 4 e + 0 1 “f: ;;;:}::::‘;1 12:71.. 322'; '7 1 . -l.22e+02 .' :; ;:.. ~1.799r02 “2.36e+02 '%.1“.j..__:;5';‘:_..._... . -2.94e+02 ' 13326:6312}???'5" -3.5]e+02 ' F ~4.08e+02 . ‘-~_;“’N -4.66e*02 ~75 5__e;~ *5.236*02 ' ‘ ' ... .0 i I relative-velocity Colored 6y X Velocity (m/sl (Time=4.4700e-041 Jul 16. 2004 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) Figure 93: Axial component of relative velocity vector, gas inlet port and channels 182 6.23e+02 5.66e+02 5.09e‘02 , _ 4.5113402 is 3.94e‘02 3.37e+02 2.79e'02 2.22e‘02 1.65e412 1.066v02 5.03: 0% -7.05e*00 -6.44e411 -l.22e‘02 '1.796’02 -2.36e'02 -2.94e+02 -3.51e412 '4.06e*02 ”4.66er02 -5.23e'02 relative- velocity Colored By X Velocity [WE] (Time-4.4700e-01.6 2004 NT 6.1 (2d. dp. coupled imp. lamlJ unsteady) 6.23e*02 5.6613412 5.09e412 4.51e412 3.94am: 3.3712412 2.7913412 2.22e~02 1.85e412 1.086412 5.03e+01 -7.05e+00 -6.44e*01 4.2212702 '1.796’02 -2.36e~02 -2.94e+02 '3.516‘02 -4.08e412 “1.6812412 -5.23e+02 relative- velocity Colored By X Velocity FIm/s) [Time= 4. 470 0e 16 2004 UT FL UENT 6.1 (.2d dp. coflu4pled imp. 1am. unsteady) Figure 94: Axial component of relative velocity, lower and upper comers of gas inlet port 183 6.23e+02 5.66e*02 5.09e412 16:1 4.516‘3? i:- ’.i 3.94et02 3.37e+02 2796*02 2.22e*02 l.65e*03 l.066+02 5.03e*Ul -7.05e+00 -6.44e*01 -l.22e41? -l.7Qer02 -2.36e+02 -2.94e*02 -3.51e*02 -4.08e+02 -4.66e*02 ‘5.23e412 1'. 7‘. ‘1' l . relative-velocity Colored By X Velocity (m/s) [Time=4.4700e-04) Jul 16. 2004 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) Figure 95: Axial component of relative velocity, air outlet port 6.23e+02 IQ; . 5.66e+02 5.09e412 . 4.51e+02 3.9412402 3.37e412 2.79et02 2.22e412 1.65-$02 1,089+02 5.03e*01 -7.05e410 -6.44e+01 -l.22e+02 -1.79e+02 -2.36e*02 -2.94e+02 -3.51e+02 -4.08e+02 ”1.6613417 ~14?“ I. "E‘:' -5.23e+02 5"“. 1"; T relative-velocity Colored By X Velocity lm/s) (Time=4.4700e-041 Jul 18. 2004 FLUENT 6.1 (2d. dp. coupled imp. lam. unsteady) Figure 96: Axial component of relative velocity, gas outlet port 184 6.2313412l . J 5.66e+02 6.0.9902 4.5163412 3.94e+02 3.37e+02 2.7919412 2.2263412 1.65e4l2 1.06e412 5.03e+01 -7.05e*00 -6.44e‘01 -l.22e+02 -1.7Qe+02 -2.366+02 -2.94e+02 “3.5let02 -4.08e*02 '4.669+02 '5.23e+02 ..., ‘. relative-velocity Colored By X Velocity [m/s) (T1me=4.4700e-04) Jul 18. 2004 FLUENT 6.1 (2d. dp. coupled imp. larn. unsteady) Figure 97: Axial component of relative velocity, air inlet port l5:“) l 1.-..FLJ “‘1'" .- . Grid Jul 19. 2004 FLUENT 6.1 (2d. dp. segregated. lam) Figure 98: Mesh for multi cycles 185 o ":. 'n CHAPTER 6: INNOVATIVE WAVE ROTOR DESIGNS AND APPLICATIONS The MSU wave rotor group has initiated studies to evaluate the possible benefits of utilizing wave rotor technology in several thermal cycle applications. Besides the efforts described so far, the team has also investigated and developed several innovative conceptual designs, which will be briefly discussed in this chapter. 6.1 Radial Wave Rotor Concept As described in the wave rotor history of Chapter 2, for various wave rotor applications several different configurations of mainly axial-flow wave rotors have been studied so far. The four-port version with straight channels has been used most widely. However, pure scavenging is a challenging task in axial-flow configurations. In gas turbine applications, neither the TF configurations nor the RF configurations can often achieve a full scavenging process, as discussed in Chapters 3 and 5. An innovative design taking advantage of centrifugal forces can improve the scavenging process, as described in the following. 6.1.1 Radial-F low Wave Rotor with Straight Channels Here, the radial-flow wave rotor concept (wave disc) is introduced employing a flow in the radial and circumferential directions. This can substantially improve the scavenging process by using centrifugal forces. Figure 99 shows schematically a simple radial-flow wave rotor with straight channels and a constant rectangular cross-sectional area. Similar to axial-flow wave rotors, if the driven flow enters and leaves at the inner radius and the driver flow enters and leaves at the outer radius, such a configuration can be referred to as a reverse-flow (RF) design. For a through-flow (TF) configuration, most 186 .t)". PM, o w '..I-' ".r .' .0 - ‘. e I“: of the driven flow travels through the full length of the channel and leaves at the outer end. Figure 99: Radial-flow wave rotor with straight channels To verify the general idea of wave disc operation, where our knowledge is still very limited, the commonly available software package FLUENT has been used to simulate the flow field inside a RF wave disk. Here, using FLUENT it has been shown that the novel concept of wave discs and their particular problems can be investigated by 2D models. The model considered here has 60 straight channels with an inner radius of 0.15 m and outer radius of 0.3 m, rotating clockwise at 8330 rpm as shown in Figure 100. The results presented below have been provided by a collaboration between the MSU wave rotor group and the wave rotor team at Warsaw University in Poland. Two sets of four ports, as in a conventional four-port wave rotor, are used. It is assumed that the fresh air enters the channels with temperature a of 300 K at a pressure of 105 Pa. It is compressed by a high-pressure, high-temperature gas entering at 1000 K and 2.105 Pa. The gas expands and leaves at the lower value of 105 Pa. 187 pressure air inflow Compressed Figure 100: Reverse-flow wave disk with straight channels From top to bottom, Figure 101 shows contours of local pressure, local temperature, and velocity. Due to the symmetry, only half of the solution is shown. Relatively uniform regions shown before and after the four ports indicate that the combination of diameters, speed, and port arrangement is not optimized, because nearly no changes are seen in these regions. Only the pressure contours show a certain radial stratification in these regions, indicating the action of centrifugal forces. They also clearly show the effects of compression and expansion waves and how the low-pressure region is created by supporting the ingestion of fresh air. The temperature contours clearly show the flow casing especially for the burned gas which is a typical feature of RF configurations. Furthermore, the contours of pressure and velocity show effects of gap leakage, especially on the left side before the high pressure gas port. 188 Figure 10]: Wave disc with straight channels: contour plots oflocal pressure (top), local temperature (middle), and velocity (bottom) at timc:8.6c-04 s 189 6.1.2 Radial-F low Wave Rotor with Curved Channels As Figure 102 suggests, the channels alternatively may be curved and varied in cross- sectional area. Compared to straight wall channels, curved channels provide a greater length for the same disc diameter, which can be important to obtain certain wave travel times for tuning. With curved channels also the angle against the radius can be changed freely. Furthermore, curved channels may be more effective for self-propelling and work extraction in the case of a wave turbine or work input for additional compression, analogous to the principle of turbomachines. Figure 102: Radial-flow wave rotor with curved channels Figure 103 is the configuration used in the numerical simulations presented here. The disc radii and number of channels are the same as for the wave disc with straight channels, as explained before. The rotational speed now is 8300 rpm. While the temperature boundary conditions for the ports also are the same as before, now two different high pressure levels are used. In the lower right comer 3.105 Pa is set for both high-pressure gas inlet and high-pressure air outlet ports while 2.10S Pa for these ports in the upper left comer, both parts being independent. 190 Figure 103: Reverse-flow wave disk with curved channels The results are presented in Figure 104, showing that compression and expansion are generally working. The static pressure contours show a radial pressure stratification in the regions where both ends of the channels are closed as discussed in Figure 101. The complete temperature contours now show a similar casing like in Figure 101, but with deeper gas penetration and an unexpected carry on of expanding hot gas stretches after the exhaust gas port opens. The penetration is obviously deeper at the right side where the high-pressure level is 3.105 Pa. More results and observations in the simulation of rather complex time-dependent flow phenomena that occur in radial wave rotors are not presented. The reader is referred to Ref. [238] for a complete discussion. 191 Figure 104: Wave disc with curved channels: contour plots of local pressure (top), local temperature (middle), and velocity (bottom), at timc=1.73e-3 s 192 6.1.3 Radial-F low Wave Rotor Concept The wave disc described above can be also stacked together as shown in Figure 105. This way a modular construction is possible that can be adapted for designs with different mass flow rates. Furthermore, similar to the known two-row axial Comprex, the channels are subdivided, which can allow for acoustic noise reduction. Such a wave disc stack can be used in the same way as a single disc wave rotor for different applications like refrigeration, gas turbine topping and supercharging of IC engines. Figure 105: Stack of wave discs with straight channels Stacked wave discs provide the unique opportunity to place a radial flow turbomachinery at the periphery of the axis of the wave disc stack with an angle such that the turbomachine impeller periphery interfaces all active discs of the stack. In this way, a peripheral continuous outflow from turbomachine impeller and inflow to the disks is possible without any additional ducting, collector, volute, diffuser or nozzle between the turbomachine impeller and the wave rotor. Thus, ducting losses are eliminated, resulting in a higher efficiency of the assembly. Figure 106 shows a configuration in which a radial compressor is placed inside a wave disc stack. Only the inner plate is placed between the impeller and the inner wave rotor. Due to the angle between the axes of the impeller and 193 wave rotor, the end plate between both can be spherical for minimum thickness (ducting length). This also allows switching on and off outer discs by varying the angle between impeller and wave rotor axis. The port opening can be a continuous oblique slot that interfaces with the impeller periphery. Since the end plates are stationary they can form one part with the housing of the turbo impeller as shown in Figure 106 for the outer impeller shroud and axial duct. The shape at the outer diameter of the wave rotor stack is generated by the shape at the inner diameter, the channel length, inclination and timing of each disc. Still if the outer shape is similar to the inner shape of the wave disc stack, the timing on each disc is different and is determined by the circumferential distance from one port to the other at the inner diameter , as shown schematically in Figure 106. Figure 106: Stacked radial wave discs and radial compressor In gas turbine applications preferably the turbo-compressor impeller is placed inside the wave rotor. Such a design eliminates the need for a diffuser which has been replaced by a more effective shock deceleration process [1, 2] in the wave disc channels. Using an outward-flow turbine, the turbine could be placed at the outer diameter with its axis also set at an angle to the wave rotor axis but rotating around the wave rotor axis with respect to the compressor axis, allowing a certain time between opening the channels at the inner and outer diameters. Such a configuration might be too challenging and would require 194 separate shafts for compressor and turbine. To avoid a gear box, their coupling could be achieved electrically via generator and motor. Figure 107 shows a simpler configuration with a direct shaft coupling compressor and turbine as in a gas turbine. This requires a flow collector from the wave rotor outer end plate and certain ducting that directs the flow to the turbine as shown schematically in Figure 108. The configurations shown here uses an internal combustion wave rotor that allows for outward flow only in the wave rotor. If a conventional external combustor is used than an additional port opening is necessary for the burned gases leaving the combustor and the high pressure air entering the combustor. External ducting may then be eliminated by having combustion in the pressure exchange channels. Figure 107: Cut~view of a radial wave rotor topping a gas turbine Figure 109 shows an exploded View of a radial wave rotor topping a gas turbine. The outer end plate is shown with an oblique slot as it would also be suitable for a peripherical outer radial outflow turbine. However, for an external turbine as shown here, the slot of the outer end plate can have any form that will adapt most conveniently to the 195 outlet opening time. In the figures sketched here, a turbine volute is used to distribute the flow around the turbine. The exhaust gas leaves the turbine axially. The reader is referred to Ref. [239] for additional discussions about the radial wave rotor concept. Fresh air Fresh air intake Intake wave rotor port/plate Internal combustion radial wave rotor Compressor Driving shaft Turbine volute Turbine Expanded gases exhaust Expanded burned gases Figure 108: Flow through an internal combustion radial wave rotor topping a gas turbine 196 9' I \ 5. compressor inlet port 4. radial compressor / 3. Internal end plate 2. external end plate 1.3tacked radial wave / rotors 6. turbine volute / 7. turbine turbine exit port Figure 109: Component parts of a radial wave rotor topping a gas turbine 6.2 Condensing Wave Rotor Wave rotor technology has the potential to enhance performance and reduce the size and cost of refrigeration cycles using water (R718) as a refrigerant. While R718 refrigeration systems have shown attractive features compared to commonly used refrigeration system, these units often suffer from the expensive and bulky multi-stage turbo-compressors which are crucial units for the R718 chiller technology. Utilizing the wave rotor in such water cycles appears to be a potentially promising solution. While possibility of integrating four-port wave rotors in R718 cycles has been discussed in a previous study [240], the attention is on utilizing a three-port wave rotor in 197 R718 units. Utilizing a three-port wave rotor, know as the condensing wave rotor, appears more promising because it combines the function of a compressor stage and the condenser in one compact unit. Both pressure rise and condensation occur inside the wave rotor channels as described below.4 6.2.1 How Does a Condensing Wave Rotor Work? Figure 110 depicts schematically a R718 cycle with a direct condenser and a direct evaporator. Figure 111 shows a schematic of a R718 cycle using a three-port condensing wave rotor. A comparison between these two cycles indicates that the wave rotor has substituted three subsystems: the condenser, one compressor stage, and the intercooler (now shown). A schematic design“ of a condensing wave rotor is depicted in Figure 112. In this innovative design, condensation of vapor occurs inside the wave rotor channels as depicted in Figure 113 and Figure 114. Figure 113 schematically shows a model for the condensation process inside a channel of a three-port condensing wave rotor. Figure 114 shows a corresponding schematic wave and phase-change diagram. 4 Materials presented in this section have been accepted for publication in 2005 ASME Journal of Engineering for Gas Turbines and Power. 198 Wm. 1 Condenser — Wcr+ Wcz Condensing wave rotor Figure ll 1: Schematic of a R718 cycle enhanced by a three-port condensing wave rotor substituting for the condenser and for one compressor stage 199 MP water (3) '0 End plate Channel I End plate Vapor collector ’ . V ‘ Q3 . Condensing Wave rotor ' .,- ‘ (drum) I HP water (6) Figure 112: Schematic of a three-port condensing wave rotor Contact surface Normal shock wave Figure 113: Regions modeled for each compression and condensation i axial length I Rotation i 1111111101111111111111111111 \1 ( uninti surface lufldlILIIJ-ill'lllflmflllllfll‘m 111m11111m11111uw ”13051101111111: Shack mm.- 200 Following the points introduced in Figure 111, coming from the turbo-compressor (2), the superheated vapor flows continuously through a vapor collector (shown in Figure 112) to the inlet port of the wave rotor located at one of the two stationary end plates. By rotating the wave rotor between the two end plates, the wave rotor channels are opened to the port and filled with the incoming superheated vapor. The region (a) in Figure 113 and Figure 114 is the state after the filling process is completed. After further rotation, the channels meet the second-inlet port (6) through which the high-pressure low-temperature water (e) comes in and is exposed to the low-pressure high-temperature superheated vapor in region (a). Due to the sudden pressure drop (from p6 to p;), all the heat cannot be contained in the incoming water as sensible heat and the heat surplus is transformed into latent heat of vaporization. This is the so called flash evaporation or flashing phenomenon [241, 242]. Therefore, one portion of the incoming water suddenly vaporizes (c) and the remaining part cools down (d). The frontal area of the saturated vapor (c) generated by the flash evaporation is called the contact interface and acts like a fast moving piston. It causes a shock wave triggered from the leading edge of the inlet port traveling through the superheated low-pressure vapor which exists inside the channel (a). The shock wave travels with supersonic speed (Vshm) faster than the contact interface (14mm...) Therefore, the trajectory of the shock wave (solid line in Fig. 6) has a smaller slope than the incoming water and the contact interface of the generated vapor (dashed line). Behind the moving shock wave (b) the temperature is increased from T 2 to T 2' and the pressure is increased from p; to p2'=p3 due to the shock compression. The latter is a design decision similar to a tuning condition. With it, the pressure at the inlet port (p6) is set to an appropriate value that generates the pressure ratio p6 /pz required to 201 trigger the desired shock wave. The superheated vapor will be condensed at pressure p3 . This shows that the fluid in its liquid state serves as a “work capacitor” storing pump work to release it during its expansion in the wave rotor channels for the simultaneous vapor compression. Therefore, in the enhanced system the pump in the cooling water cycle not only has to provide the work necessary to overcome the pressure loss in the heat rejecter cycle (wPL) but also the work necessary for the shock wave compression in the wave rotor channels (war). The pressure behind the shock wave (b) is imposed on the vapor generated by the flash evaporation (c). It is the pressure at the water surface and the equilibrium pressure at which the evaporation decays p(c)=p(b)=p3. Hence, both generated vapor and the cooled water obtain the saturation temperature T 3=Tm(p 3). Due to the contact of the superheated compressed vapor (b) with the cold incoming water, the superheated vapor is desuperheated and its heat is transferer (I) to the incoming water. This continues until the equilibrium temperature T3 is achieved in region (b) and the superheated vapor is changed to saturated vapor. Subsequently, the incoming water compresses the saturated vapor further and condenses it, while the latent heat is transferred to the incoming water (g). The water, which is nearly a fully condensed two- phase vapor with a typical quality of 0.005, is scavenged through the only outlet of the wave rotor (3). The scavenging process may be supported by gravity and pump power. The schematic pressure-enthalpy (p-h) and temperature-entropy (T-s) diagrams of both the baseline and the wave—rotor-enhanced cycle are depicted in Figure 115 and Figure 116, respectively. 202 ------ l-2b-3b-5b-6b-3b-4b baseline cycle 1-2-2' -3-5-6-3-8 wave rotor enhanced cycle / Pressure Enthalpy Figure 115: Schematic p-h diagram of a R718 baseline cycle and a wave rotor enhanced cycle ------ 1-2b-3b-4b baseline cycle l-2-2’ -3-5-6-3-8 wave rotor enhanced cycle i: 3 N 6 D- 6 3 [— 6 ...Y. i 3 1135/ '1 “'33" 4:8 Entropy Figure 116: Schematic T-s diagram of a R718 baseline cycle (cooling water cycle not shown) and a wave rotor enhanced cycle Both cycles start at the outlet of the evaporator l, where the vapor is saturated. State 21, represents the compressor outlet of the baseline cycle whereas state 2 is the compressor outlet of the wave—rotor-enhanced cycle that allows using a compressor with a lower pressure ratio. State 2' is an intermediate state inside the wave rotor channels that corresponds to the flow properties in region (b) right after the shock wave. The slope between points 2 and 2' is greater than that between points 1 and 21, because the shock compression typically occurs with a higher efficiency. Still inside the wave rotor channel, the superheated vapor is desuperheated to the equilibrium temperature T 3 (2'—>3). State 3 is actually much closer to the liquid region than shown in Figure 115 and Figure 116 because the mass flow rate of the cooling water cycle is much greater than that of the refrigerant cycle. Knowing this, it becomes clear that the distances between points 3, 5, and 6 are exaggerated in both diagrams. The expansion process (6—>3) releases the energy consumed by the compression process of the vapor (2—>2’) all within the wave rotor channels. Coming from the only outlet port of the wave rotor (3), the flow diverges. The small fraction used as refrigerant is directed to the expansion valve and is expanded in a constant enthalpy process (3—> 4), while most of the flow out of the wave rotor goes into the heat rejecter (cooling tower or similar) where it cools (3—> 5). Afterwards the pressure is again increased (5——>6) by the pump, providing the energy for the vapor compression in the wave rotor (wwc) and compensating for the pressure loss in the heat rejecter and associated piping (wPL). 6.2.2 Performance Evaluation Using a thermodynamic model described in Ref. [243, 244], a performance map of the enhanced cycle is obtained as shown in Figure 117. For each point on this plot, the mass flow ratio between the cooling water cycle and the chilled water cycle is considered as a fixed value, e.g. here, K =m6 /m2=125. Reference [245] describes how increasing the mass flow ratio above 200 appears ineffective for the COPgam (COP increase of enhanced cycle relative to the base cycle divided by the COP of the baseline cycle) and the gradient 204 of the COPgam increase above K =125 is not significant. According to Figure 117, each point on this plot shows the maximum COPgam that can be obtained by the optimum choice of PRW for a given evaporator temperature and a temperature-lift meaning that for each evaporation temperature and temperature-lift combination, PRW is varied within a certain range (1