You are on page 1of 350
16.Unified: Thermodynamics and Propulsion 2006-2007 Prof. Z. S. Spakovszky Prof. I. A. Waitz LSTEX editing by D. Quattrochi Notes by E. M. Greitzer, Z. S. Spakovszky, 1. A. Waitz Origins ‘These notes represent the combination of two previously separate courses: “16.Unified: ‘Ther- modynamies and Propulsion,”and “16,05: ‘Thermal Energy,” which evolved independently from 1996 to 2006. Acknowledgement from 16.05 Notes Preparation of these notes has benefited greatly from the expertise of a number of individuals, and we are pleased to acknowledge this help. ‘The successive graduate and undergraduate ‘Teaching Assistants in this core department course, Jessica ‘Townsend, Vincent Blateau, Isabel Pauwels, David Milanes, Doug Quattrochi, Nick Chan and Julia Thrower provided ideas, corrected errors, inserted “Muddy Points,” supplied the index, and in general, created a much more readable document. We are greatly indebted to Doug Quattrochi for hi tremendous support and invaluable help throughout this course, and for typesetting and editing the document in KTEX. Any errors that remain, or lack of readability, are thus the sole responsibility of the authors. We also appreciate the work of Diana Park and Robin Palazzolo, who helped editing the graphics. Finally, we are grateful to have had the opportunity to discuss some of the material with Professor Frank Marble of Caltech, whose understanding, insight, and ability to describe thermofluids concepts provide a model of how to address important technical problems. A note on bracketed references Sections of these notes contain bracketed references to other texts, primarily © VN: H.C. Van Ness, Understanding Thermodynamics, Dover Publications, 1983; and © SBKVW: R. E. Sonntag, C. Borgnakke, and G. J. Van Wylen, Fundamentals of Ther- ‘modynamics, John Wiley Publishers, 1998. Contents 1 THE FIRST LAW OF THERMODYNAMICS 1 Introduction to Thermodynamics 1.1 What it’s All About z 12. Definitions and Fundamental Kdeas of Thermodynamics 1.2.1 The Continuum Model 1.2.2 The Concept of a “System se The Concept of a “State” ‘The Concept of “Equilibrium’ The Concept of a “Process” Quasi-Equilibrium Processes .2.7 Equations of state erty 1.3 Changing the State of a System with Heat and Work... . 13.1 Heat... pevectenmumme say 1.3.2 Zeroth Law of Thermodynamics 1.3.3 Work 1.3.4 Work vs. Heat ~ which 1.4 Muddiest Points on Chapter 1 2 The First Law of Thermodynamics 2.1 First Law of Thermodynamics 2.2 Corollaries of the First Law . 2.3. Example Applications of the First Law; enthalpy inbatic, steady, throttling of a gas 23.2 Quasi-Statie Expansion of a Gas 2.3.3 ‘Transient filling of a tank 23.4 The First Law in Terms of Enthalpy 24 Specific Heats . . 2.4.1 Specific Heats of an Ideal Gas 2.4.2 Reversible adiabatic processes for an ideal gas... . . « 2.5 Control volume form of the system laws 2.5.1 Conservation of mass 2.5.2 Conservation of energy . 2.5.3 Stagnation Temperature and Stagnation Enthalpy 2.5.4 Example Applications of the Steady Flow Energy Equation 25 25 25 26 2 30 aL 33 36 36 38 38. n 26 Muddiest Points on Chapter 2 ‘The First Law Applied to Engineering Cycles 3.1, Some Properties of Engineering Cycles; Work and Bificiency . . . 3.2 Generalized Representation of Thermodynamic Cycles . 3.3 The Camot Cycle 3.4 Refrigerators and Heat Pumps 3.5 The Internal combustion engine (Otto Cycle). 3.5.1 Efficiency of an ideal Otto cycle eas 3.5.2 Engine work, rate of work per unit enthalpy flax... 3.6. Diesel Cycle 3.7. Brayton Cycle. rsa 3.7.1 Work and Eficiency . . - 3.72 Gas Turbine Technology and Thermodynamics... . 3.7.3 Brayton Cyele for Jet Propulsion: the Ideal Ramjet 3.74 MIT Cogenerator . mae H RETR EROS 3.8 Muddiest points on Chapter THE SECOND LAW OF THERMODYNAMICS Background to the Second Law of Thermodynamics 4.1 Reversibility and Irreversibility in Natural Processes. « 4.2 Difference between Free and Isothermal Expansions 4.3 Features of reversible processes 4.4 Muddiest Points on Chapter 4 ‘The Second Law of Thermodynamics 5.1 Concept and Statements of the Second Law... . 6... 5.2 Axiomatic Statements of the Laws of Thermodynamics... . . 5.2.1 Introduction 5.22 Zeroth Law . 5.2.3 First Law 5.24 Second Law 5.2.5 Reversible Processes eee Combined First and Second Law Expressions 53 5.4 Entropy Changes in an Ideal Gas . 5.5. Caleulation of Entropy Change in Some Basie Processes 5.6 Muddiest Points on Chapter Applications of the Second Law 6.1 Limitations on the Work that Can be Supplied by a Heat 6.2 The Thermodynamic ‘Temperature Scale 6.3. Representation of Thermodynamic Processes in T-s coordinates . 64 ‘ton Cycle in T-s Coordinates GAL Net work por unit mass flow in a Brayton c} ngine le CONTENTS oT 59 60. 02 63, 05, or or 09 a a 2 7 81 82 82 89 CONTENTS 6.5 Irreversibility, Entropy Changes, and “Lost Work” 6.6 Entropy and Unavailable Energy 6.7 Examples of Lost Work in Engineering Processes 6.8 Some Overall Comments on Entropy 6.8.1 Entropy : 682 Reversible and Irreversible Processes . 6.83 Examples of Reversible and Irreversible Processes 6.9 Effect of Departures from Ideal Behavior... 6.0... ees 6.9.1 Parameters reflecting design choices 6.9.2 Parameters reflecting the ability to design and execute efficient com- ponents 6.10 Muddiest Points on Chapter € 7 Entropy on the Microscopic Scale itropy Change in Mixing of Two Ideal Gases 7.2. Microscopie and Macroscopic Descriptions of a System 7.3 A Statistical Defin 7.5 Numerical Example: the Equilibrium 7.6 Summary and Conclusions 8 Power Cycles with Two-Phase Media 8.1 Behavior of Two-Phase Systems 8.2 Work and Heat Transfer wi 83. The Carnot Cycle as a Two-Phase Power Cycle . 83.1 Example: Carnot steam eycle . 84 The Clausits-Clapeyron Equation 8.5 Rankine Power Cycles. « 8.6. Enhancements of Rankine Cycles 8.7 Combined Cycles for Power Production 8.8 Some Overall Comments on Thermodynamic Cycles 8.9 Muddiest Points on Chapter & h Two-Phase Medi Il PROPULSION 9 Introduction to Propulsion 9.1 Goal: Create a Force to Propel a Vehicle... . 9.2. Performance parameters . 9.3 Propulsion isa systems endeavor 10 Integral Momentum Theorem 10.1 Au Expression of Newton's 2" Law . 102 Application of the Integral Momentum Equation to Rockets 10.3 Application of the Momentum Equation to an Aircraft Engine 124 = 128 131 1d ML ut 142 143 M45, 195 197 - 197 198 198 201 - 201 204 204 6 CONTENTS 11 Aireraft Engine Performance 11.1 Overall Efficiency - 112 Thermal and Propulsive Efficiency z 11.3 Iinplications of propulsive efficiency for engine design . . 11.4 Other expressions for efficiency ~ J4p and SFC 11.5 Trends in thermal and propulsive efficiency 116 Performance of Jet Engines. 11.6.1 Notation and station numbering 11.6.2 Ideal Assumptions 11.6.3. Ideal Ramjet 11.64 Turbojet Engine 11.7 Performance of Propellers . sua 6 11.7.1 Overview of propeller performance . . . 11.72 Application of the Integral Momentum Theorem to Propellere 11.7.3 Actuator Disk Theory 11.74 Dimensional Analysis. . 11.8 Muddiest points on Chapter 11. 12 Aircraft Performance 12.1 Vehicle Drag 12.2 Power Required . oe 1233 Aireraft Range: the Breguet Range Equation 12.3.1 Relation of overall efficiency, zp, and thermal effi 12.3.2 ‘The Propulsion Energy Conversion Chain 124 Aireraft Endurance 12.5 Climbing Flight : vee 13 Rocket Performance 13.1 Thrust and Specific Impulse for Rockets oo... ee 13.2 The Rocket Equation . 133 Rockot Nozzles: Connection of Flow to Geometry 13. 13.3.2 Thrust terms of nozzle geometry . . 14 Energy Exchange with Moving Blades 14.1 Introduction 14.2 Conservation of Angular Mom 14.3 The Euler Turbine Equation 14.4 Multistage Axial Compressors 165 VelociyTiisugho for an Kal Coingrenot Senge 146 Velocity Triangles for an Axial Flow Turbine Stage ‘tum 1 Quas-one-dimensional compressible How in a variable area duct... - CONTENTS IV HEAT GENERATION AND TRANSFER 15 Generating Heat: Thermochemistry 15.1 15.2 oe 15.3. Enthalpy of formation a 15.4 First Law Analysis of Reacting Systems 15.5 Adiabatic Flame Temperature . 16.5.1 Approximate solution using “average” values of specific heat 15.6.2 Solution for adiabatic flame temperature using evolutions of specific heats with temperature 15.5.3 Solution for adiabatic flame temperature using tabulated values for - 280 gas enthalpy 15.6 Muddiest points on Chapter 15 16 Conductive Heat Transfer 16.1 Heat ‘Transfer Modes ni 16.2 Introduction to Conduction . . . 16.3 Steady-State One-Dit nduction sional C 16.3.1 Example: Heat transfer through a plane slab... 16.4 Thermal Resistance 16.5 Quasi-One-Dimensional Heat Flow 16.5.1, Cylindrical Shell Spherical Shell. . 16.6 Muddiest Points on Chapter 16 nits 17 Convective Heat Transfer 17.1 The Reynolds Analogy . 17.2 Combined Conduction and Convection 17.3. Dimensionless Numbers and Analysis of Ri 17.4 Muddiest Points on Chapter 17 18 Generalized Conduction and Convection 18.1 Temperature Distributions in the Presence of Heat Sources 18.2 Heat Transfer From a Fin 18.3 Transient Heat Transfer (Convective Cooling or Heating) 18.4 Modeling Complex Phy 18.5 Heat Exchangers 185.1 Simplified Cousterflow Heat Exchanger (With U ature) 18.5.2 General Counterflow Heat Exchanger... 6.0.00. 18.5.3 Efficiency of a Counterfiow Heat Exchanger . 18.6 Muddiest Points on Chapter 18 cal Processes form Wall Temper- 280 283 283, . 284 285, 286 288, 290 290 = 293 294 297 298, - 303 307 - 307 a - 311 313, 318, - 319 320 323, 326 328, 8 CONTENTS 19 Radiation Heat Transfer 329 19.1 Ideal Radiators 2368 19.2 Kirchhoff’s Law and “Real Bodies . 19.3 Radiation Heat Transfer Between Planar Surfaces 19.3.1 Example 1: Us 19.3.2 Example 2: ‘Temperat transfer sei 19.4 Radiation Heat Transfer Between 19.4.1 Example: Concentric ¢ 19.5 Muddiest Points on Chapter 18 of a thermos bottle to reduce heat transfer sasurement error due to radiation heat V_ APPENDIX: THERMODYNAMIC PROPERTIES IN SI UNITS. 343 VI Index 345 List of Figures 1.1 The propulsion chain . 1.2 Examples of heat: en} : 13. Piston (boundary) and gas (system) 14 Boundary around electric motor (system) 1.5 Sample control volume . . 1.6 Equilibrium . . Le LT ‘Thermodynamics coordinates and isotines for an ideal gas. . 1.8 The zeroth law schematically 1.9 A closed system (dashed box) against a piston, which moves into the sur roundings 28 1.10 Work during an irreversible process . . . 28 1.11 Work in P-V coordinates 29 1.12 Simple processes 29 1.13 Quasi-stati, isothermal expansion of an ideal gas 30 1.14 A resistor heating water . 31 1.15 Pressure-temperature-volume surface for a substance that expands on freezing 32 2.4 Random mot 1n is the physical basis for internal energy... - wn OM 2.2 The change in energy of a system relates the heat added to the work done . . 35 23 The change in energy hetween two states is not path dependent. ....... 36 24 Energy is a function of state only. 0... eee ee eee eee ee 86 2.5 Adiabatic flow through a valve, a generic throttling process .......... 87 2.6 Equivalence of actual system and piston model, 37 2.7 A transient problem — filling of a tank from th 39 2.8 Control volume and system for flow through a propulsion device a7 2.9 A control volume used to track mass flows . . . aT 2.10 Schematic diagrams illustrating terms in the energy equation for a simple and a more general control volume 49 2.11 Streamlines and a stagnation region eee BL 2.12 A stationary gas turbine drawing air in from the atmosphere. . + 53 2.13 A control volume approach to the tank filling problem... .......... BA 2.14 Flow through a rocket nozzle isn ds sw ae BO 2.15 The Pratt and Whitney 4084.0... 0.0... s cece sees seseam BB 3.1 Examples of the conversion of work into heat . 60 9 10 LIST OF FIGURES Isothermal expansion 6. A generalized heat engine... . : 02 Carnot eycle . er) Work and heat transfers in a Carnot eycle between two heat reservoirs... 6 Operation of a Carnot refrigerator 65 Schematic of a domestic refrigerator «6... sss eee eee wsasho OA ‘The ideal Otto cycle . < ete wwecgepyszase Sketch of an actual Otto cycle 68 Piston and valves in a four-stroke internal combustion engine . . 69 Ideal Otto cycle thermal efficiency . 6.0. ee eee eee eee ee 70 ‘The ideal Diesel cycle. : . 7 Sketch of the jet engine components and corresponding thermodynamic states 72 Schematies of typical military gas turbine engines... .. . ene eee 15 Thermodynamic model of gas turbine engine cycle for power generation... 74 Options for operating Brayton cycle gas turbine engines 4 Gas turbine engine pressures and temperatures % Gas turbine engine pressure ratio trends « seven 18 ‘Trend of Brayton cycle thermal efficiency with compressor pressure ratio 7 Rolls-Royce high temperature technology... 6-0. ee 7 ‘Turbine blade cooling technology technology . . 78 Efficiency and work of two Brayton cycle engines 78 33 Trend of cycle work with compressor pressure ratio 80 Aeroengine core power... oe eee eee eee eee 8L 5 The ideal ramjet... 0 oe ee eee eee eee eee . 82 Brayton cycle considered as a number of elementary Carnot cycles +. 86 4.1, Flywheel in insulated enclosure at initial and final states... 00.02... 91 4.2 Bricks separated by a temperature difference 92 4.3. Free expansion i 93 4.4. Expansion against a piston = 93 4.5 Returning the free expansion to its intial condition . » -. 4.6 Work and heat exchange in the reversible isothermal compression process... 94 4.10 4a 4.12 100% conversion of work into heat... AGRwUMINcarszeaaae 98 Work and heat transfer in reversible isothermal expansion... . 9% A piston with weights on top : 97 Getting the most work out ofa system... 6. ee eee ee seenee, Heat transfer across a finite temperature difference... . . tesoams OO icolas Sadi Camot 98 ‘The Kelvin-Planck statement of the Second Law . 102 ‘The Clausius statement of the Second Law 102 Heat transfer from/to a heat reservoir - 109 Heat transfer between two reservoirs. - 110 Work from a single heat reservoir . 110 ‘The “Hot Brick” Problem A 5 ul LIST OF FIGURES 64 62 63 64 65 66 67 68 69 6.10 eat 6.12 613 614 6.15 6.16 aT 6.18 Ta 72 73 74 75 Ba 82 83 84 85 86 BT 88 89 8.10 sun gaz 813 8d B15 8.16 A Camot eycle heat engine Arrangement of heat engines to demonstrate the thermodynan seale . : Carnot eycle in T-s coordinates = ‘ Ideal Brayton cycle as composed of many elementary Carnot cycles Arbitrary cycle operating between Tinins Trnax + aio 8 98 : Brayton cycle in enthalpy-entropy (Ii-s) representation showing compressor and turbine work cone Irreversible and reversible state changes Adiabatic Throttling... . . te Static and stagnation pressures and temperatures . 5... +. Stagnation and statie states Losses reflected in changes in stagnation pressure when T; = const. Schematic of turbine and representation in h-s coordinates . Isothermal expansion with friction ‘A cycle operating irreversibly between two temperature reservoirs ic temperature Airfoil with control volume for analysis of propulsive power requirement . . . Gas turbine engine (Brayton) cycle showing effect of deverse from ideal behavior in compressor and turbine Nor-dimensional power and efficiency for a non-ideal gas turbine engine Scale diagram of non-ideal gas turbine cycle . Some allowed states of the system in the numerical example . . . Constant energy state groups . . . : ‘Transition probabilities in numerical experiment with isolated system Evolution of the probability distribution wi itropy for the system as a function of tim ‘Two-phase system in contact with constant temperature heat reservoir... . ~ 168 P-T relation for a liquid-vapor system -v diagram for two-phase system showing isotherms . - Constant pressure curves in T-v coordinates showing vapor dome “ Specific volumes at constant temperature and states within the vapor dome in a liquid-vapor system : Liquid vapor equilibrium in a two-phase mediuin Carnot eycle with two-phase medium Carnot eycle devised to test the validity of the laws of thermodynamics Clansius-Clapeyron Experimental Proof (1) - Clausius-Clapeyron Experimental Proof (2) . Rankine power cycle with two-phase working fluid Rankine cycle diagra Rankine cycle with superheating Comparison of Rankine cycle with superheating aud Caruot cycle Rankine cycle with snperheating and reheat for space power application Effect of exit pressure on Rankine cycle efficiency +. 174 . 179 - + BL - 182 - 183 = 125 125 131 - 132 132 133 - 134 136 137 139 143 146 M7 -- 16L - 162 163 164 - 164 168 169 - 169 - 170 171 1st - 185 - 186 186 187 2 LIST OF FIGURES 8.17 Effect of maximum boiler pressure on Rankine cycle efficiency 187 8.18 Gas turbine-steam combined cycle... . . . « 3 188 8.19 Schematic of combined cycle using gas turbine and steam turbine 189 8.20 Comparison of efficiency and power output of various power products 190 8.21 The vapor dome in T-s (experimental data) 193 8.22 Clausius-Clapeyron Experimental Proof (3) 194 94 al liquid propellant rocket motor pum sagsasce ee 1h 9.2 Schematies of typical military gas turbine engines . . wees 198 9.3. Schematics of a PW PT6A-65, a typical turboprop . eres 9.4 The RB211-535E4, a typical high bypass-ratio turbofan... . . exas3a 10.1 Two inertial coordinate systems, one stationary, one translating . 201 10.2 Falling blocks 202 10.3 Control volume for application of momentum theorem to a rocket. = 204 10.4 Control volume for application of momentum theorem to an aircraft en 205 10.5 Alternate control volume for application of momentum theorem to an aircraft engine . 206 11.1 Schematic of the inlet area of an engine + 208 11.2 Propulsive efficiency and specific thrust as a function of exhaust velocity . . . 209 11.3 The F-22 Raptor - 210 11.4 The Boeing 777-200 - 210 11.5 A propeller gives a rlatively small impulse (Au) to a relatively large mass flow211 11.6 An advanced, contour-rotating, unducted fan concept 212 11.7 Propulsive efficiency comparison for vatious gus turbine engine configurations 213 11.8 Trends in aireraft engine efficiency... 6.0... eee ee eee pene och 11.9 Pressure ratio trends for commercial transport engines ©... vs... 5. 5 215: 11.10 Trends in turbine inlet temperature sound a3 en 4508 eedeIA 11.11 Trends in engine bypass ratio wh 11.12Gas turbine engine station numbering, ++ 218 11.13Control volume over the burner . $223 u Dis 11.147 rust per unit mass flow and specific impulke for ideal ramjet 218 11.15Schematic with appropriate component notations added... 6... +. ++. 219 1116Performance of an ideal turbojet engine vs. fight Mach number . 24 11.17Performance of an ideal turbojet engine vs. compressor pressure ratio, fi Mach number 11.18Performance of an ideal turbojet engine ¥: inlet temperature - cee 2 BB2 11.19Cessna Skyhawk single engine propeller plane... 4... ++ geen ei838 11.20The V-22 Osprey utilizes tiltrotor technology 233 11.21Schematic of propeller... . . « eee 24 11.22Control volume for analysis of a propeller + 235 11.23Schematic of actuator disk model = 236 11.24Control volume around actuator disk. 236 11.25Typical propeller efficiency curves vs. advance ratio and blade angle 237 LIST OF FIGURES 9 Velocity triangles for an axial flow turbine stage. | propeller thrust curves ws. advance ratio and blade angle propeller power curves vs. advance ratio and blade angle . . A schematic of the forees on an aircraft in steady level flight . . . Components of vehicle drag. . ‘Typical power required curve for an aitcraft ‘The propulsion energy conversion chain fro Part 1 13 237 238, 241 - 242 Relationship between condition for maximum endurance and maximum range. 247 Force balance for an aircraft in climbing flight. ‘Typical behavior of power available as a fnction of flight velocity. Lockheed Martin F-16 performing a vertical accelerated climb. Schematic of rocket nozzle and combustion chamber Schematic for application of the momentum theorem. ‘The Saturn V rocket General form of relationship between flow area and Mach number. Control volume around compressor or turbine. Control volume for Euler Turbine Equation. . A typical multistage axial flow compressor . . ee Schematic representation of an axial flow compressor... 5... < Pressure and velocity profiles through a multi-stage axial compressor Velocity triangles for an axial compressor stage Compressor behavior . . . matic of an axial flow turbine. 1 Constant pressure combustion. . . eee 2 Specific heat as a function of temperature . . . 13 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8 16.9 16.10Cylindrieal shell geometry notati 16.11Spherical shell Schematie of adiabatic flame temperature . . Conduction heat transfer... 6. Heat transfer along a bar One-dimensional heat conduetion — ‘Temperature boundary conditions for aslab ss... sss. ‘Temperature distribution throughs a slab we Heat transfer across a composite slab (series thermal resistance) Heat transfer for a wall with dissimilar materials (parallel thermal resistance) Heat transfer through an insulated wall ‘Temperature distribution through an insulated wall 16.12Interfuce Control Volume awa 172 Turbine blade heat transfer configuration Temperature and velocity distributions near a surface - 264 - 266 = 266 267 - 273 276 277 ++ 283 28h =. 286 - + 287 - 287 288, 289) 289 290 = 291 293, 295, 297 298 u LIST OF FIGURES 173 Velocity profile near a surface 299 174 Momentum and energy exchanges in turbulent flow... 66... 2... 299 17.5 Heat exchanger configurations 302 17.6 Conducting wall with convective heat transfer. a SOS 17.7 Cylinder in a flowing fluid nase 2 305 178 Critical radius of insulation... . ee ~ 306 17.9 Effect of the Biot Number on temperature distributions... = 808 17.10 Temperature distribution in a convectively cooled eylinder for different values of Biot number 309 18.1 Slab with heat sources 311 18.2 Temperature distribution for slab with distributed heat sources 313 18.3 Geometry of heat transfer fin teen ee 314 18.4 Element of fin showing heat transfer... 2... ee 31d 18.5 The temperature distribution ina fin... 0.20.02. BIT 186 ‘Temperature variation in an object cooled by a flowing fluid 318 18.7 Voltage change in an R-C circuit... . cima ees acess a cS 18.8 Concentrie tubes heat exchangers... 06.0.0 e eee eae ett 18.9 Cross-flow heat exchangers... . . . ee cee BAL 18.10Geometry for heat transfer between two fluids acum anes 2 eee IBRD 18.11Counterflow heat exchanger... . 18.12Fluid temperature distribution along the tube with uniform wall temperature 323, 19.1 Radiation surface properties . 19.2 Emissive power of a black body at several temperatures 19.3 A cavity with a small hole (approximates a black body) . . « 19.4 A small black body inside a cavity . 19.5 Path of a photon between two gray surfaces 19.6 Schematic of a thermos wall . as 19.7 Thermocouple used to measure temperature - 19.8 Effect of radiation heat transfer on measured temperature 19.9 Shielding a thermocouple to reduce radiation heat transfer error 19.10Radiation between two bodies... . . . 19.11 Radiation between two arbitrary surfaces 19.12Radiation heat transfer for concentric eylinders or 19.13View Factors for three-dimensional geometries . . 19.14View factor for aligned parallel rectangles . . parallel disk... . . 19.16View factor for perpendicular rectangles with a common edge List of Tables 11.1 Turbojet Mission Parameters 223, 11.2 Propeller Parameters and their Units 228 16.1 Thermal conductivity at room temperature for some metals and non-metals . 285 16.2 Utility of plane slab approximation 293 19.1 ‘Total emittances for different surfaces 339) 16 LIST OF TABLES Part I THE FIRST LAW OF THERMODYNAMICS Chapter 1 Introduction to Thermodynamics [VN Chapter 1] 1.1 What it’s All About ico a "Thermodynamics is a sei J, more importantly, ant engineering tool used to describe processes that involve changes in temperature, transformation of energy, and the relation- ships between heat and work. It can be regarded as a generalization of an enormous body of empirical evidence. It is extremely general: there are no hypotheses made concerning th structure and type of matter that we deal with. It is used to describe the performance of propulsion systems, power generation systems, and refrigerators, aud to describe fluid flow, combustion, and many other phenomena. The focus of thermodynamics in aerospace engineering is on the production of work, often in the form of kinetic energy (for example in the exhaust of a jet engine) or shaft power, from different sources of heat. For the most part the heat will be the result of combustion processes, but this is not always the case. The course content can be viewed in terms of a “propulsion chain” as shown in Figure 1.1, where we see a progression from an energy source ‘to useful propulsive work (thrust power of a jet engine). In terms of the different blocks, Parts I and I ire mainly about how to progress from the second block to the third, Part IIT takes us from the third to the fourth, and a chapter in Part IV takes us from the first to the second. We will start with the progression from heat to work, examples of which are given in Figure 1.2 Energy source 7 ‘Mechanical ‘Useful propulsive| chemical (combustion work, work (thrust nuclear, ete process) >| electric power. Figure 1.1: The propulsion chain "Alternatively, most thermodynamie principles and ean be derived from kinetic theory 19 20 CHAPTER 1. INTRODUCTION TO THERMODYNAMICS ‘Combustion Heat - Mechanical Work Solar Heat et | Heat ghee: Electrical Energy Nuclear Heat N. ‘Waste Heat (a) Converting heat to useful work <*> ae ~ mcr (b) Propulsion system (6) Power gener ‘Mechanical Work Electrical Energy — [———— | Heat (©) Converting work to useful heat Figure 1.2: Examples of heat engines Practice Questions 1. Describe the energy exchange processes in (blank) (fil in the blank, e.g. a nuclear power plant, a refrigerator, a jet engine). 2. Given that energy is conserved, where does the fuel+oxidizer energy that is used to power an airplane go? 3. Describe the energy exchange processes necessary to use electricity from a nuclear power plant to remove heat from the food in a refrigerator. 4. Describe the energy exchange processes necessary for natural gas to be used to provide electricity for the lights in the room you are in 1.2. Definitions and Fundamental Ideas of Thermodynamics As with all sciences, thermodynamics is concerned with the mathematical modeling of the real world. In order that the mathematical deductions are consistent, we need some precise definitions of the basic concepts. The following is a discussion of some of the concepts we will need. Several of these will be further amplified in the lectures and in other handouts. If you heed additional information or examples concerning these topics, they are described clearly and in-depth in (SB&VW). They are also covered, although in a tess detailed manner, in Chapters 1 and 2 of the book by Van Ness. INS AND FUNDAMENTAL IDEAS OF THERMODYNAMICS 21 1.2.1 The Continuum Model Matter may be described at a molecular (or microscopic) level using the techniques of statis- tical mechanics and kinetic theory. For engineering purposes, however, we want “averaged” information, ile., a macroscopic, not a microscopic, description. ‘There are two reasons for this. First, a microseopie description of an engineering device may produce too much information to manage. For example, 1 mm* of air at standard temperature and pressure contains 10" molecules (VW, $ & B:2.2), each of which has a position and a veloci engineering applications involve more than 10% molecules. Second, and more importantly, microscopic positions and velocit nerally not useful for determining how n systems will act or react unless, for instance, their total effect is integrated. We therefore neglect the fact that real substances are composed of discrete molecules and model matter from the start as a smoothed-out continuum. The information we have about a continmum represents the microscopic information averaged over a volume. Classical thermodynamics is concerned only with continua, 1.2.2 The Concept of a “System” A thermodynamic system is a quantity of matter of fixed identity, around which we ean draw a boundary (see Figure 1.3 for an example). The boundaries may be fixed or moveable. Work or heat can be transferred across the system boundary. Everything outside the boundary is the surroundings. When working with devices such as engines it is often useful to define the system to be an identifiable volume with flow in and ont. ‘This is termed a eontrol volume. An example is shown in Figure 1.5, A closed system is a special class of system with boundaries that matter cannot cross. Hence the principle of the conservation of mass is automatically satisfied whenever we employ a closed system analysis. This type of system is sometimes termed a control mass. Figure 1.3: Piston (boundary) and gas (system) 1.2.3. The Concept of a “State” ‘The thermodynamic state of a system is defined by specifying values of a set of measurable properties sufficient to determine all other properties, For fluid systems, ical properties are pressure, volume and temperature. More complex systems may require the specification 22 CHAPTER 1. INTRODUCTION TO THERMODYNAMICS System boundary Figure 1 pt —> 1 J 1] complex | | iL eeess Hy na too-aT Figure 1.5: Sample control volume of more unusual properties. As an example, the state of an electric battery requires the specification of the amount of electrie charge it contains. Properties may be extensive or intensive. Extensive properties are additive. ‘Thus, if the system is divided into a number of sub-systems, the value of the property for the whole system is equal to the sum of the values for the parts. Volume is an extensive property. Intensive properties do not depend on the quantity of matter present. Temperature and pressure are intensive properties. Specific properties are extensive properti letters. For example: per unit mass and are denoted by lower ease specific volume = V/m =v. Specific properties are intensive because they do not depend on the mass of the system. The properties of a simple system are uniform throughout. In general, however, the properties of a system can vary from point to point. We can usually analyze a general system by subdividing it (either conceptually or in practice) into a number of simple systems in each of which the properties are assumed to be uniforin, Tt is important to note that properties describe states only when the system is in eqni- librium. Muddy Points Specific properties (MP 1.1) INS AND FUNDAMENTAL IDEAS OF THERMODYNAMICS What is the difference between extensive and intensive properties? (MP 1.2) 1.2.4 The Concept of “Equilibrium” ‘The state of a system in which properties have definite, unchanged values as long as external conditions are unchanged is called an equilibrium state, Pe Mase] A as | gas | rttt 1 7 ab TK oat Copper Patton Mg+P,A=pa Pressure, P Overtime, T) = Tp (6) Mechanica! Equilibrium (b) "Thermal Equiibrium Figure 1.6: Equilibrium sm in thermodynamic equilibrium satisfies: 1. mechanical equilibrium (no unbalanced forces) 2, thermal equilibrium (no temperature differences) 3. chemical equilibrium, 1.2.5 The Concept of a “Process” If the state of a system changes, then it is undergoing a process. ‘The succession of states through which the system passes defines the path of the process. If, at the end of the process, the properties have returned to their original values, the system has undergone a ceyelic process or a eyele. Note that even if a system has returned to its original state and completed a cycle, the state of the surroundings may have changed. 1.2.6 Quasi-Equilibrium Processes We are often interested in charting thermodynamic processes between states on thermody- namie coordinates. Recall from the end of Section 1.2.3, however, that properties define a state only when a system is in equilibrium. If a process involves finite, unbalanced force the system can pass through non-equilibrium states, which we cannot treat. An extremely useful idealization, however, is that only “infinitesimal” unbalanced forces exist, so that the process can be viewed as taking place in a series of “quasi-equilibrium” states. (The term quasi can be taken to mean “as if” you will see it used in a mumber of contexts such as quasi-one-dimensional, quasi-steady, etc.) For this to be true the process must be slow in relation to the time needed for the system to come to equilibrium internally. For a gas at conditions of interest to us, a given molecule can undergo roughly 10! molecular collisions By CHAPTER 1. INTRODUCTION TO ‘THERMODYNAMICS per second, so that, if ten collisions are needed to come to equilibrium, the equilibration time is on the order of 10-® seconds. ‘This is generally much shorter than the time scales associated with the bulk properties of the flow (say the time needed for a fluid particle to move some significant fraction of the length of the device of interest). Over a large range of Parameters, therefore, it is a very good approximation to view the thermodynamic processes ‘sting of such a suce ession of equilibrium states, which we can chart. [VW, S& B: 4) ‘The figures below demonstrate the use of thermodynamics coordinates to plot isolines, lines along which a property is constant. ‘They include constant temperature lines, or isotherms, on a p-v diagram, constant volume lines, or isochors on a ‘T-p diagram, and constant pressure lines, or isobars, on a ‘7-v diagram for an ideal gas. Real substances may have phase changes (water to water vapor, or water to ice, for example), which we can also plot on thermodynamic coordinates. We will see such phase changes plotted and used for liquid-vapor power generation cycles in Chapter 8. A preview is given in Figure 1.15 at the end of this chapter. ang temperature gag Pressure AF Ba os am 10 isis Specie Volume ia) i Torporatiu (Kein) (@) pv diagram () PT diogram rereasing pressure 0 io 1a ‘Spec Vol (3g) (©) Pov diagram Figure 1.7: Thermodynamics coordinates and isolines for an ideal gas 1.3, CHANGING THE S (ATE OF A SYSTEM WITH HEAT AND WORK 1.2.7 Equations of state It is an experimental fact that. two properties are needed to define the state of any pure substance in equilibrium or undergoing a steady or quasi-steady process. [VW, S & B: 3.1, 3.3]. Thus for a simple compressible gas like air, P=P(v,T), or v=u(P,T), or T=T(P.v), where v is the volume per unit mass, 1/p. In words, if we know v and 7 we know P, ete. Any of these is equivalent to an equation (P,v, 7) =0, whieh is known as an equation of state. The equation of state for an ideal gas, which is a very good approximation to real gases at conditions that are typically of interest for aerospace application:?, is Po=RT, where J is the volume per mol of gas and is the “Universal Gas Constant,” 8.31 kJ/kmol-K. obtained if we A form of this equation which is more useful in fluid flow problem divide by the molecular weight, M: Pu=RY, or P=pRT where R is R/.M, which has a different value for different gases due to the different molecular weights. For air at room conditions, R = 0.287 kJ /kg-K. 1.3 Changing the State of a System with Heat and Work ‘Changes in the state of a system are produced by interactions with the environment through heat and work, which are two different modes of energy transfer. During these interactions, ibrium (a static or quasi-static process) tem properties to one-another to be val wecessary for the equations that relate sy 1.3.1 Heat Heat is energy transferred due to temperature differences only. 1. Heat transfer can alter system states; 2. Bodies don’t “contain” heat; heat is identified as it comes across system boundaries; The amount of heat needed to go from one state to another is path dependent; 4. Adiabatic processes are ones in which no heat is transferred. 2p << Poon, and T > 2Te up U0 about Apone, [VW, $ & B: 3.4} 26 CHAPTER 1. INTRODUCTION TO THERMODYNAMICS 1.3.2 Zeroth Law of Thermodynamics With the material we have discussed so far, we are now in a position to describe the Zeroth Law. Like the other laws of thermodynamics we will see, the Zeroth Law is based on observation. We start with two such observations: for a sufficiently 1. Iftwo bodies are in contact through a thermally-conducting boundat long time, no further observable changes take place; thermal equilibrium is said to prevail. 2, Two systems which are individually in thermal equilibrium with a third are in thermal equilibrium with each ot! called temperature. ; all three systems have the same value of the property These closely connected ideas of temperature and thermal equilibrium are expressed formally in the “Zeroth Law of Thermody1 Zeroth Law: ‘There exists for every thermodynamic system in equilibrium a prop- erty called temperature, Equality of temperature is a necessary and sufficient condition for thermal equilibrium. The Zevoth Law thus defines a property (temperature) and describes its behavior Note that this law is true regardless of how we measure the property temperature. (Other relationships we work with will typically require an absolute seale, so in these notes we use either the Kelvin K = 273.15+°C or Rankine R = 459.9+° F scales. Tempera be discussed further in Section 6.2.) The zeroth law is depicted schematical —- <> ifT\-T and then ;Q)= 0 Figuro 1.8: The zoroth law schematically Section 5.2 for the otler laws and exedit For this form, 1.3, CHANGING THE STATE OF A SYSTEM WITH HEAT AND WORK a 1.3.3. Work [VW, S & B: 4.1-4.6] Section 1.8.1 stated that heat is a way of changing the energy of a system by virtue of a ‘temperature difference only. Any other means for changing the energy of a system is called work. We can have push-pull work (e.g. in a piston-cylinder, lifting a weight), electric and magnetic work (e.g. an electric motor), chemical work, surface tension work, elastic work, etc. In defining work, we focus on the effects that the system (e.g. an engine) has on its igs. Thus we define work as being positive when the system does work on the surroundings (energy leaves the system). If work is done on the system (energy added to the system), the work is negative. Consider a simple compressible substance, for example, a gas (the system), exerting a force on the surroundings via a piston, which moves through some distance, [ (Figure 1.9), ‘The work done on the surroundings, Won suer. 8 Won sure, = Foree on surr. x dl Force on sur. Won sure, = ——z——— (Area x di) fron * (Area xd) | NTN (molecular motion) (molecular motion) Figure 2.1: Random motion is the physical basis for internal energy 2. The change in energy of a system is equal to the difference between the heat added 10 the system and the work done by the system, AE=Q-W_ (units are Joules, J), (1) where # is the energy of the system, Qs the heat input to the system, and W is the work done by the E=U (thermal energy) + Exinetic + Epotentiat + + (a) Like the Zeroth Law, the First Law describes the behavior of the new property [VW, S& B: Chapter 5]. (b) The equation can also be written on a per unit mass basis Ac g—w (units are J/kg). (c) Im many situations the potential energy, kinetic energy, and chemical energy of the system are constant or not important. Then AE = AU, and AU=Q-W or Au=q-w. (@) Note that Q and 1 are not functions of state, but U, which arises from molecular motion (see above), depends only on the state of the system; U does not depend stem got to that state. We therefore have the striking result that: AU is independent of path even though Q and IV’ are not! Sometimes this difference is emphasized by writing the First Law in differential form, dU=6Q-6W or du whore the symbol “6” is used to denote that these are not exact differentials but rather are dependent on path. bq — dw, (22) 21, FIRST LAW OF THERMODYNAMICS 35 (€) Note that the signs are important: ‘© Qs defined to be positive if itis transferred 0 the system; thns the numerical value we substitute for Q will be positive if heat is transferred fo the system from the surroundings, and negative if heat is transferred from the system to the surroundings. (VW, $ & B: 4.7-4.8] ‘* IV is defined to be positive if it is done by the system (see Section 1.3); thus the numerical value we substitute for I will be positive if the system is doing work, and negative if work is being done on the system. [VW, S& B: 4.1-4.4] (f) For quasi-static processes we can substitute W = paysdV, dU =6Q-plV or dw=6q~pdv ‘To give an example of where the first law is applied, consider the device shown in Fig- ure 2.2. We heat a gas, it expands against a weight, some force (pressure times area) applied over a distance, and work is done. The change in energy of the system supplies the connection between the heat added and work done. We will spend most of the course dealing with various applications of the first law — in one form or another. Pressure _ Heat (Q) Figure 2.2: The change in energy of a system relates the heat added to the work done The form of the first law we have given here is sometimes called the “control mass” form, because it is well suited to dealing with systems of a fixed mass. We will see in Section 2.5 that this form can be written for a control volume with mass flow in and mass flow out (like fa jet engine for example). We will call this the “control volume” form of the first law [VW, S&B: 5.85.12]. Muddy Points What are the conventions for work and heat in the first law? (MP 2.1) When does + U? (MP 2.2) 36 CHAPTER 2, THE FIRST LAW OF THERMODYNAMICS 2.2 Corollaries of the First Law 1. Work done in any adiabatic (Q = 0) process is a function of state, We ean write the first law, setting the heat transfer term equal to zero, as 0 AU =-W. (2.3) Since AU depends only on the state change, now W ean be found as a function of the state change. Figure 2.3: The change in energy between two states is not path dependent. 2. For a cyclic process heat and work transfers are numerically equal a Figure 2.4: Since energy is a function of state only, any process that returns a system to its original state leaves its energy unchanged. Grint therefore and Q=W oor fia- faw 2.3 Example Applications of the First Law to motivate the use of a property called “enthalpy” IVW, § & Bi 54 2.3.1 Adiabatic, steady, throttling of a gas (flow through a valve or other restriction) Figure 2.5 shows the configuration of interest. We wish to know the relation between prop- erties upstream of the valve, denoted by “1” and those downstream, denoted by “2” 23, EXAMPLE APPLICATIONS OF THE FIRST LAW; ENTHALPY 37. Figure 2.5: Adiabatic flow through a valve, a generic throttling process pistons Figure 2.6: Equivalence of actual system and piston model ‘To analyze this situation, we can define the system (choosing the appropriate system i often a critical element in effective problem solving) as a mnit mass of gas in the following two states. Initially the gas is upstream of the valve and just through the valve. In the final state the gas is downstream of the valve plis just hefore the valve. ‘The figures on the left of Figure 2.6 show the actual configuration just described. In terms of the system ert ‘the same pressure that the external fluid exerts, as indicated schematically on the right side of Figure 2.6. The process is adiabatic, with changes in potential energy and kinetic energy assumed to be negligible. The first law for the system is therefore behavior, however, we could replace the fluid external to the system by pistons which ex au =-W. ‘The work done by the 3} W = PoVa~ PiMi- Use of the first law leads to Unt PaVe = Ui + PAV. In words, the initial and final states of the system have the same value of the quantity U-+ PV. For the case examined, since we are dealing with a unit mass, the initial and final states of the system have the sate value of u + Pv. We define this quantity as the “enthalpy,” usually denoted by H, H=U+PVv. 38 CHAPTER 2. THE FIRST LAW OF THERMODYNAMICS In terms of the specific quantities, the enthalpy per unit mass is h=utPv ut Pip. It is a function of the state of the system. H has units of Joules, and h has units of Joules per kilogram, The utility and physical significance of enthalpy will become clearer as we work with more flow problems. For now, you may wish to think of it as follows (Levenspiel, 1996) ‘When you evaluate the energy of an object of volume V, you have to remember that the object had to push the surroundings out of the way to make room for itself. With pressure p on the object, the work required to make a place for itself is pV. This is so with any object or system, and this work may not be negligible. (The force of one atmosphere pressure on one square meter is equivalent to the force of a mass of about 10 tons.) ‘Thus the total energy of a body is its internal energy plus the extra energy it is credited with by having a volume V at pressure p. We call this total energy the enthalpy, HT. Muddy Points ‘When is enthalpy the same in initial and final states? (MP 2.3) 2.3.2 Quasi-Static Expansion of a Gas Consider a quasi-static process of constant pressure expansion. We can write the first law in terms of the states before and after the expansion as Q=(U2-U1)+W. and writing the work in terms of system properties, = W2—U;) + pVe-Vi) since 1 = pa =p. By grouping terms we can write the heat input in terms of the enthalpy change of the system: Q = (U2 + pV2) — (Oi + Vi) = Ha-Hy 2.3.3 Transient filling of a tank Another example of a flow process, this time for an unsteady flow, is the transient process of filling a tank, initially evacuated, from a surrounding atmosphere, which is at’ a pressure Py and a temperature To. The configuration is shown in Figure 2.7. Ata given time, the valve at the tank inlet is opened and the outside air rushes in. The inflow stops when the pressure inside is equal to the pressure outside. ‘The tank is insulated, so there is no heat transfer to the atmosphere. What is the final temperature of the gas in the tank? This time we take the system to be all the gas that enters the tank. The initial state has the system completely ontside the tank, and the final state has the system completely inside . EXAMPLE APPLICATIONS OF THE FIRST LAW; ENTHALPY 39, Po.Te | Vacuum Valve Figure 2.7: A transient problem ~ filling of a tank from the atmosphere the tank. The kinetic energy initially and in the final state is negligible, as is the change i potential energy, so the first law again takes the form AU =-wW. Work is done on the system, of magnitude PyVo, where Vj is the initial volume of the system, 0 AU = PoVo. In terms of quantities per unit mass (AU = mAu, Vo = mo, where m is the mass of the system), Au = tana = wi = Porn ‘The final value of the internal energy is To. ‘The final temperature is thus roughly 200°F hotter than the outside air! It may be helpful to recap what we used to solve this problem. There were basically four steps: 1. Definition of th 2. Use of the first law 3. Equating the work to a “PdV” term 4. Assuming the fluid to be a perfect gas with constant specific heats. 40 CHAPTER 2. THE FIRST LAW OF THERMODYNAMIC: A message that can be taken from both of these examples (as well as from a large number of other more complex situations, is that the quantity h = u+Pv occurs naturally in problems of fluid flow. Because the combination appears so frequently, it is not only defined but also tabulated as a function of temperature and pressure for a number of working fluids. Muddy Points In the filling of a tank, why (physically) is the final temperature in the tank higher than the initial temperature? (MP 2.4) 2.3.4 The First Law in Terms of Enthalpy We start with the first law in differential form and substitute pdV for dW by assuming a quasi-static or reversible process: a du 6Q- SW (true for any process, neglecting AKE and APE) 6Q—pdV (true for any quasi-static process, no AK E or APE) ‘The definition of enthalpy, H usp, can be differentiated (applying the chain rule to the pV term) to produce dH =dU + pdV +Vap. rst. Law, we obtain Substituting the dV above for the dU in the dH =6Q-6W+pdV+Vdp (valid for any process) dH =6Q+Vdp (valid for any quasi-static process). 2.4 Specific Heats: the relation between temperature change and heat [VW, S& B: 5.6] How much does a given amount of heat transfer change the temperature of a substance? It depends on the substance. In general Q=CAT, (2.4) where C is a constant that depends on the substance, We can determine the constant for any substance if we know how much heat is transferred. Since heat is path dependent, however, we must specify the process, ie., the path, to find C. ‘Two useful processes are constant pressure and constant volume, so we will consider these cach in turn. We will call the specific heat at constant pressure Cp, and that at constant volume Oy, oF ¢p and ¢y per unit mass. 24, SPECIFIC HEATS a 1. The Specific Heat at Constant Volume Remember that if we specify any two properties of the system, then the state of the system is fully specified. In other words we can write u = u(L,v), u = u(p,v) or u=u(p,T). [VW, 8 & B: 5.7] Consider the form u = u(,v), and use the chain rule to write how w changes with respect to T and v: ou ou tu = (Se a) ae. 2.5) =( ‘control volume A control volume used to track mass flows 2.5.2 Conservation of energy ‘The first law of thermodynamics can be written as a rate equation: ab ‘, anew where sto the system = tin (82 Q= Jim, (2) rate of total heat trai fon (8 W = tim (S- rate of total work done by the system, amo dt 4s. CHAPTER 2. THE FIRST LAW OF THERMODYNAMICS To derive the first law as a rate equation for a control volume we proceed as with the mass conservation equation. ‘The physical idea is that any rate of change of energy in the control volume must be caused by the rates of energy flow into or out of the volume. Th heat transfer and the work are already included and the only other contribution must be associated with the mass flow in and out, which carries energy with it. Figure 2.10 shows two schematics of t idea. The desired form of the equation will be (riseigrtnee) = Gases.) — (maceo") + (me share fo") — (gore) « The fluid that enters or leaves has an amount of energy per unit mass given by +2/2+g2, where ¢ is the fluid velocity relative to some coordinate system, and we have neglected chemical energy. In addition, whenever fluid enters or leaves a control volume there is a work term associated with the entry or exit. We saw this in Section 2.3, example 1, and ‘the present derivation is essentially an application of the ideas presented there. Flow exiting at station “e” must push back the surrounding fluid, doing work on it. Flow entering the volume at station “i” is pushed on by, and receives work from the surrounding air. The rate of flow work at exit is given by the product of the pressure times the exit area times the rate at which the external flow is “pushed back.” The latter, however, is equal to the volume per ‘unit mass times the rate of mass flow. Put another way, in a time dé, the work done on the surroundings by the flow at the exit station is dWoow = mde. ‘The net rate of flow work is Wow = Pevetite piviiiie Including all possible energy flows (heat, shaft work, shear work, piston work, etc.), the first law can then be written as: 4y Bey = 3 Geet Went +3 Wanear +0 Wyant Waow +o th (« + 2 +s) where D> includes the sign associated with the energy flow. If heat is added or work is done on the system then the sign is positive, if work or heat are extracted from the system then the sign is negative. NOTE: this is consistent with AE = Q—W, where IV is the work done by the system on the environment, thus work is flowing out of the system. We can then combine the specific internal energy term, tu, in ¢ and the specific flow work term, pr, to make the enthalpy appear: ‘Total energy associated with mass flow: etpaute/2+g2+ pu hte /2rgz —> | pei PROL VOLUME FORM OF THE SYSTEM LAWS: Figure 2.10: Schematic diagrams illustr a more general control volume Waar (a) Simple (b) More General 49 terms in the energy equation for a simple and 50 CHAPTER 2. THE FIRST LAW OF THERMODYNAMICS ‘Thus, the first law can be written as: S DY Fa = Qe + YL Wares + YO Waroar + YO Whiston + 2 rhe ( + + «) For most of the applications in this course, there will be no shear work and no piston work. Hence, the first law for a control volume will be most often used as: dE dt 2 é dey — Wenate + 124 (" + 2 +0) — tie (n+ z +0) (2.10) Note how our use of enthalpy has simplified the rate of work term. In writing the control volume form of the equation we have assumed only one entering and one leaving stream, but this could be generalized to any number of inlet and exit streams. In the special case of a steady-state flow, £ @ and thy = the = th. Applying this to Equation 2.10 produces a form of the “Steady Flow Energy Equation” (SFEB), , =i|(esdees)-(neZrm)). ean which has units of Joules per second. We could also divide by the mass flow to produce net @ ned = (ter Foo) (0 Eo), which has units of Joules per second per kilogram. For problems of interest in aerospace applications the velocities are high and the term that is associated with changes in the elevation is small. From now on, we will neglect the gz terms unless explicitly stated. Jew — Wev er — Muddy Points What is shaft work? (MP 2.5) ‘What distinguishes shaft work from other works? (MP 2.6) Definition of a control volume (MP 2.7) 2.5.8 Stagnation Temperature and Stagnation Enthalpy Suppose that our steady flow control volume is a set of streamlines dese to the nose of a blunt object, as in Figure 2.11 ‘The streamlines are stationary in space, so there is no external work done on the fluid as it flows. If there is also no heat transferred to the flow (adiabatic), then the steady flow energy equation becomes ng the flow up 25. CONTROL VOLUME FORM OF THE SYSTEM LAWS Et Figure 2.11: Streamlines and a stagnation region; a control volume can be drawn between the dashed streamlines and points 1 and 2 ‘The quantity that is conserved is defined as the stagnation temperature, 2 T+ in or M? using a = />RT, where M = c/a is the Mach nnmber®. ‘The stagnation temperature is the temperature that the fluid would reach if it were brought to zero speed by a steady adiabatic process with no external work. Note that for any steady, adiabatic flow with no external work, the stagnation ‘temperature is constant. It is also convenient to define the stagnation enthalpy, 2 é Ne = OT +5 which allows us to write the Steady Flow Energy Equation in a simpler form as 12 ~ tha = he ha Note that for a quasi-static adiabatie process z-(@)’ 50 we can write "The Mac mumber, 2, is the ratio of the flow speed, ¢, to the speed of sound, a. You will learn more about this quantity in fluids, but itis interesting to see that MM? measures the ratio of the kinetic energy of the gas to its therm energy. 52 CHAPTER 2. THE FIRST LAW OF THERMODYNAMICS and define the relationship between stagnation pressure and Static pressure as Ba (rs sr)’ P where, the stagnation pressure is the pressure that the fluid would reach if it were brought to zero speed, via a steady, adiabatic, quasi-static process with no external work. Frame dependence of stagnation quantities An area of common confusion is the frame dependence of stagnation quantities. The stagna- tion temperature and stagnation pressure are the conditions the fluid would reach if it were brought to zero speed relative to some reference frame, vin a steady adiabatic process with no external work (for stagnation temperature) or a steady, adiabatic, reversible process with no external work (for stagnation pressure). Depending on the speed of the reference frame the stagnation quantities will take on different values. For example, consider a high speed reentry vehicle traveling through the still atmosphere, which is at temperature, 7. Let's place our reference frame on the vehicle and stagnate a fluid particle on the nose of the vehicle (carrying it along with the vehicle and thus essentially giving it kinetic energy). The stagnation temperature of the air in the vehicle frame is Natta, where c is the vehicle speed. The temperature the skin reaches (to first approximation) is the stagnation temperature and depends on the speed of the vehicle. Since re-entry vehicles travel fast, the skin temperature is much hotter than the atmospheric temperature. ‘The atmospheric temperature, T,, is not frame dependent, but the stagnation temperature, T;, is, ‘The confusion comes about because 7’ is usually referred to as the static temperature, In common language this has a similar meaning as “stagnation,” but in fluid mechanics and thermodynamics static is used to label the thermodynamic properties of the gas (p, T, ete.), and these are not frame dependent. Thus in our re-entry vehicle example, looking at the still atmosphere from the vehicle frame we see a stagnation temperature hotter than the atmospheric (static) temperature. If we look at the same still atmosphere from a stationary frame, the stagnation temperature the same as the statie temperature Example For the case shown below, a jet engine is sitting motionless on the ground prior to take- off. Air is entrained into the engine by the compressor. ‘The inlet can be assumed to be frictionless and adiabatic. Considering the state of the gas within the inlet, prior to passage into the compressor, as state (1), and working in the reference frame of the motiontess airplane: 25. TROL VOLUME FORM OF THE SYSTEM LAWS Atmosphere: Taxon Pan Inlet Exhaustiet, Figure 2.12: A stationary gas turbine drawing air in from the atmosphere 1. Is Ta greater than, Jess than, or equal to Tom? ‘The moving the same speed as the reference frame (the motionless airplane). ‘The steady flow energy equation tells us that if there is no heat or shaft work (the case for our adiabatic inlet) the stagnation enthalpy (and thns stagnation temperature for constant Cy) remains unchanged. Thns Ty = Team stagnation temperature of the atmosphere, Tyan, is equal to Tatm since it is 2. . Is T; greater than, less than, or equal to Tytm? IE Tir = Taam then Ty < Taam since the flow is moving at station 1 and therefore some of the total energy is composed of kinetic energy (at the expense of internal energy, thus lowering 7}) . Is pa greater than, less than, or equal to pat Equal, by the same argument as 1 » . Is pi greater than, less than, or equal to Pac Less than, by the same arguinent as 2. Steady Flow Energy Equation in terms of Stagnation Enthalpy ‘The form of the “Steady Flow Energy Equation” (SFEE) that we will most commonly use is Equation 2.11 written in terms of stagnation quantities, and neglecting chemical and potential energies, Steady Flow Energy Equation: Qev — Wanate = ti (Fine ~ has) ‘The steady flow energy equation finds much use in the analysis of power and propulsion devices and other finid machinery. Note the prominent role of enthalpy. Muddy Points What is the difference between enthalpy and stagnation enthalpy? (MP 2.8) 2.5.4 Example Applications of the Steady Flow Energy Equation [VW, S& B: 6.4] ET CHAPTER 2, THE FIRST LAW OF THERMODYNAMICS = 1 control surface | con ‘ | volume ‘m_(qass flow) =! 1 L. Figure 2.13: A control volume approach to the tank filling problem ‘Tank Filling Using what we have just learned we can attack the tank filling problem solved in Section from an alternate point of view using the control volume form of the first law. In this problem the shaft work is zero, and the heat transfer, kinetic energy changes, and potential energy changes are neglected. In addition there is no exit mass flow. ‘The control volume form of the first law is therefore aw at = rnghi. "The equation of mass conservation is Combining we have * dU _ dm a a ie to the final time (the incoming enthalpy is hy = ho as before. hi Integrating from the ini using U = mu gives the result upnat constant) and Flow through a rocket nozzle A liquid bi-propellant rocket consists of a thrust chamber and mn forcing the liquid propellants into the chamber where they react, converting chemical energy to thermal energy. Once the rocket is operating we can assume that all of the flow processes are steady, 50 it is appropriate to use the steady flow energy equation, Also, for now we will assume that the gas behaves as a perfect gas with constant specific heats, though in general this is a poor approximation. There is no external work, and we assume that the flow is adiabatic. We define our control volume as going between location c, in the chamber, and location ¢, at the exit, and then write the First Law ax male and some means for doe Wace = hte he — which becomes hye = hee 2.5. CONTROL VOLUME FORM OF THE SYSTEM LAWS fie combustion 2% mb . oxide fe hot igh pressure Tow vlonty g38| Figure 2.14: Flow through a rocket nozzle or ‘Therefore T, and pe, the conditions in the combustion chamber, are set by propellants, and pe is the external statie pressure. Power to drive a gas turbine compressor Consider for example the PW4084 pictured in Figure 2.15. ‘The engine is designed to produce about 84,000 Ibs of thrust at takeoff. The engine is a two-spool design. The fan and low pressure compressor are driven by the low pressure turbine. The high pressure compressor is driven by the high pressure turbine. We wish to find the total shaft work required to drive the compression system. ‘my = total pressure ratio across the fan wld otal pressure ratio across the fan + compressor M5 10 kes 20 kg/s 56 CHAPTER 2. THE FIRST LAW OF THERMODYNAMICS Figure 2.15: The Pratt and Whitney 4084 (drawing courtesy of Pratt and Whitney) We define our control volume to encompass the compression system, from the front of the fan to the back of the fan and high pressure compressor, with the shaft cutting through the Heat transfer from the gas streams is negligible, so we write back side of the control v the First Law (steady flow energy equation) p = W, = rile — ha) ; and the core stream, ¢ For this problem we must consider two streams, the fan stream, Wy = ring Ahes + tin Aline = rngepATig + tee AT ie We obtain the temperature change by assuming that the compression process is quasi and adiabatic, T € ) i Am then Ta) _ 3! . = =a,” =1.1> ATi fn = 30K GE). _— = Foe = 3.0 ATi core = 600 K 26, MUDDIES ‘T POINTS ON CHAPTER ?? a7. Substituting these values into the expression for the first law above, along with estimates of ¢p» we obtain Ww, = 10 ky/s x 80 K x 1008 J/! 91 x 10° J/s 1 Megawatts negative sign implies work done on the fluid +120 kg/s x 600 K x 1008 J/kg Note that 1 Hp = 745 watts. If'a car engine has ~ 110 Hp = 8.2 10* watts, then the power needed to drive compressor is equivalent to 1,110 automobile engines. All of this power is generated by the low pressure and high pressure turbines. 2.6 Muddiest Points on Chapter 2 MP 2.1 What are the conventions for work and heat in the first law? Heat is positive if it is given to the system. Work is positive if it is done by the system. MP 2.2 When does E> U? We deal with changes in energy. When the changes in the other types of energy (Kinetic, po- tential, strain, etc.) can be neglected compared to the changes in thermal energy, then it is a good approximation 10 use AU as representing the total energy change. MP 2.3 When is enthalpy the same in initial and final states? Initial and final stagnation enthalpy is the same if the flow is steady and if there is no net shaft work plus heat transfer. Ifthe change in kinetic energy is negligible, the initial and final enthalpy is the same. The tank problem is unsteady so the initial and final enthalpies are not the same. See the discussion of the steady flow energy equation in notes, Section 2.5 MP 24 In the filling of a tank, why (physically) is the final temperature in the tank higher than the initial temperature? Work is done on the system, which in this problem is the mass of gas that is pushed into the tank, MP 25 What is shaft work? Tam not sure how best to answer, but it appears that the difficulty people are having might be associated with being able to know when one can say that shaft work occurs. There are several features of a process that produces (or absorbs) shaft work. First of all the view taken of the process is one of control volume, rather than control mass (see the discussion of control volumes in Chapter I or in IAW). Second, there needs to be a shaft or equivalent device (a moving belt, a row of blades) that can be identified as the work carrier. Third, the shafi work is work over and above the “flow work” that is done by (or received by) the streams that exit and enter the control volume. MP 2.6 What distinguishes shaft work from other works? The term shaft work arises in using a control volume approach. As we have defined it, “shaft work” is all work over and above work associated with the “flow work” (the work done by pressure forces). Generally this means work done by rotating machinery, which is carried by a shaft from the control volume to the ouside world. There could also be work over and above the pressure force 58 CHAPTER 2. THE FIRST LAW OF THERMODYNAMICS work done by shear stresses at the boundaries of the control volume, but this is seldom important if the control boundary is normal to the flow direction. If'we consider a system (a mass of fixed identity, say a blob of gas) flowing through some dev’ neglecting the effects of raising or lowering the blob the only mode of work would be the work to compress the blob. This would be true even ifthe blob were flowing through a turbine or compressor. (in doing this we are focusing on the same material as it undergoes the unsteady compression or expansion processes in the device, rather than looking at a control volume, through which mass passes The question about shaft work and non shaft work has been asked several times. I am not sure how best fo answer, but it appears that the difficulty people are having might be associated with being able to know when one can say that shaft work occurs. There are several features of a process that produces (or absorbs) shaft work. First of all, the view taken of the process is one of control volume, rather than control mass (see the discussion of control volumes in Section 2.5 or in IAW). Second, there needs 10 be a shaft or equivalent device (a moving belt, a row of blades) that can be identified as the work carrier. Third, the shaft work is work over and above the flow work that is done by (or received by) the streams that exit and enter the control volume, MP 2.7 Definition of a control volume. A control volume is an enclosure that separates a quantity of matter from the surroundings or environment. The enclosure does not necessarily have to consist of a solid boundary like the walls of a vessel. Inis only necessary that the enclosure forms a closed surface and that its properties are defined everywhere. An enclosure may transmit heat or be a heat insulator. It may be deformable and thus capable of transmitting work 10 the system. It may also be capable of transmitting mass. MP 28 What is the difference between enthalpy and stagnation enthalpy? The enthalpy of a gas is defined ash = w+ pv = «+ p/p, and represents both the internal energy of that state and the flow work done on the gas to get it at that pressure and density. The stagnation enthalpy of a gas is defined as hy = h-+ &/2 and accounts for both the enthalpy and the Kinetic energy of the gas at that state. Chapter 3 The First Law Applied to Engineering Cycles [VW, S & B: Chapter 9, 11.8, 11.9, 11.40, 11.11, 11.12, 11.13, 11.14] This chapter is devoted to describing the fundamentals of how various heat engines work. A heat engine is a device that uses heat to produce work, or uses work to move around heat. Refrigerators, internal combustion (automobile) engines, and jet engines are all types of heat engines. We will model theso heat engines as thermodynamic cycles and apply the First Law of Thermodynamics to estimate thermal efficiency of pressures and temperatures at various points in the cycle. This is called ideal eycle analy: ‘The estimates we obtain from the analysis represent the best achievable performance that reality, the performance of these systems will be and work output as a function may be obtained from a heat engine. somewhat less than the estimates obtained from ideal cycle analysis ~ you will learn how to make more realistic estimates late Our analyses will use the “air-standard eyele,” w behavior. Specifically, we make the following simplifications: h is an approximation to actual eyele # Air is the working fluid (the presence of fuel and combustion products is neglected)! © Combustion is represented by heat transfer from an external heat source, ‘The eycle is ‘completed’ by heat transfer to the surroundings, Alll processes are internally reversible (described more fully in Chapter 5), # Airis a perfect gas with constant specific heats. Muddy Points How can we idealize fuel addition as heat addition? (MP 3.1) Hin general this is a good assumption since in typical combustion applications the fuel aceounts for only about 5% of the mass of the working fui. 60. CHAPTER 3. THE FIRS > + oar R Block on rough surface Viscous liquid Resistive heating ’ LAW APPLIED TO ENGINEERING CYCLES Figure 3.1: Examples of the conversion of work into heat 3.1 Some Properties of Engineering Cycles; Work and Effi- ciency As preparation for our discussion of eycles (and as a foreshadowing of the second law), we examine two types of processes that concern interactions between heat and work. ‘The first of these represents the conversion of work into heat. ‘The second, which is much more useful, concerns the conversion of heat into work. ‘The question we will pose is how efficient can this conversion be in the two enses. ‘Three examples of the first process are given in Figure 3.1. ‘The first is the pulling of a block on @ rough horizontal surface by a force which moves through some distance, Frietion resists the pulling. After the force has moved through the distance, it is removed. The block then has no kinetic energy and the sane potential energy it had when the force started to act. If we measured the temperature of the block and the surface we would find that it was higher than when we started. (High temperatures can be reached if the velocities of pulling are high; this is the basis of welding.) The work done to move the block has been converted totally to heat The second example concerns the stirring of a viscous liquid. There is work associated with the torque exerted on the shaft turning through an angle. When the stirring stops, the fluid comes to rest and there is (again) no change in kinetic or potential energy from the initial state. The fluid and the paddle wheels will be found to be hotter than when we started, however. ‘The final example is the passage of a current through a resistance. This is a case of electrical work being converted to heat, indeed it models operation of an electrical heater. All the examples in Figure 3.1 have 100% conversion of work into heat. This 100% conversion could go on without limit as long as work were supplied. Is this true for the conversion of heat into work? To answer the last question, we need to have some basis for judging whether work is done in a given process. One way to do this is to ask whether we can construct a way that the process coukd result in the raising of a weight in a gravitational field. If so, we can say “Work has been done.” Tt may sometimes be difficult to make the link between a complicated thermodynamic process and the simple raising of a weight, but this is a rigorous test for the existence of work. One example of a process in which heat is converted to work is the isothermal (constant 3.1, SOME PROPERTIES OF ENGINEERING CYCLES; WORK AND EFFICIENCYG1 Br Pam }—+ Work received, W t | ° Figure 3.2: Isothermal expansion temperature) expansion of an ideal gas, as sketched in Figure 3.2. The system is the gas inside the chamber. As the gas expands, the piston does work on some external device. For an ideal gas, the internal energy is a function of temperature only, so that if the temperature is constant for some process the internal energy change is zero. ‘To keep the temperature constant during the expansion, heat must be supplied. Because AU =0, the first law takes the form Q = W. This is a process that has 100% conversion of heat into work. ‘The work exerted by the system is given by rp Work = | Pav where 1 and 2 denote the two states at the beginning and end of the process. ‘The equation of state for an ideal gas is P=NRIJV, with the number of moles of the gas contained in the chapber. Using the equation of state, the expression for work can be written as 2 a avjV =NRTIW () - GA) Work during an isothermal expansion = NRT For an isothermal process, PY = constant, so that Pi/P2 = Va/Vi written in terms of the pressures at the beginning and end as Work during an isothermal expansion (2) ‘The lowest. pressure to which we can expand and still receive work from the system is atmospheric pressure. Below this, we would have to do work on the system to pull the piston ont further. ‘There is thus a bound on the amount of work that can be obtained in the isothermal expansion; we cannot continue indefinitely. For a power or propulsion system, however, we would like a source of continuous power, in other words a device that would give power or propulsion as long as fuel was added to it. ‘To do this, we need a series of processes where the system does not progress through a one-way transition from an initial state to a different final state, but rather cycles back to the initial state. What is looked for is in fact a thermodynamic cycle for the system. We define several quantities for a cycle: 62 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES © Quis the heat absorbed by the system. © Qp is the heat rejected by the system. ‘© W is the net work done by the system. ‘The cycle returns to its initinl state, so the overall energy change, AU, is zero. ‘The net work done by the system is related to the magnitudes of the heat absorbed and the heat rejected by W =Net work =Qa—-Qr. ‘The thermal efficiency of the cycle is the ratio of the work done to the heat absorbed. (Efficiencies are often usefully portrayed as “What you get” versus “What you pay for.” Here what we get is work and what we pay for is heat, or rather the fuel that generates the heat.) In terms of the heat absorbed and rejected, the thermal efficiency is Work done _ @a~@r _,_ Qe a = thermal offieney = FP absorbed = Qa a) The thermal efficiency can only be 100% (complete conversion of hent into work) if Qn = 0; basic question is what is the maximum thermal efficiency for any arbitrary cycle? ‘We examine this for several cases, including the Camot eycle and the Brayton (or Joule) cycle, which is a model for the power cycle in a jet engine. 3.2 Generalized Representation of Thermodynamic Cycles 8 & B: 6.1) Ivw Before we examine individual heat engines, note that all heat engines can be represented generally as a transfer of heat from a high temperature reservoir to a device, whieh does work on the surroundings, followed by a rejection of heat from that device to a low temperature reservoi Figure 3.3: A generalized heat engine 3.3, THE CARNOT CYCLE Fi af tcl | » ,| , Comte we He st} emer Imtngaand Reser Figure 3.4: Carnot eycle — thermodynamic diagram on left and schematic of the different stages in the eycle for a system composed of an ideal gas on the right 3.3 The Carnot Cycle A Camot cycle is shown in Figure 3.4. It has four processes. ‘There are two adiabatic reversible legs and two isothermal reversible legs. We can construct a Carnot cycle with many different systems, but the concepts can be shown using a familiar working fluid, the ideal gas. ‘The system can be regarded as a chamber enclosed by a piston and filled with this ideal gas. ‘The four processes in the Carnot eycle are: 1. The system is at temperature Ty at state a. It is bronght in contact with a heat reservoir, which is just a liquid or solid mass of large enough extent such that its temperature does not change appreciably when some amount of heat is transferred to the system. In other words, the heat reservoir is a constant temperature source (or receiver) of heat. The system then undergoes an isothermal expansion from a to 6, with heat absorbed Qo. At state b, the system is thermally insulated (removed from contact with the heat reservoir) and then let expand to ¢. During this expansion the temperature decreases to T;. The heat exchanged during this part of the cycle, Qe = 0.) 3. At state ¢ the system is brought in contact with a heat reservoir at temperature Ty. It is then compressed to state d, rejecting heat Q: in the process. 4, Finally, the system is compressed adiabatically back to the initial state a. ‘The heat exchange Qua ‘The thermal efficiency of the cycle is given by the definition Qe a ate (a) In this equation, there is a sign convention implied. The quantities Qa, Qx as defined are the magnitudes of the heat absorbed and rejected. The quantities Q1, Q2 on the other hand are defined with reference to heat received by the system. In this example, the former ” 4 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES Figure 3.5: Work and heat transfers in a Carnot eyele between two heat reservoirs is negative and the latter is positive. The heat absorbed and rejected by the system takes place during isothermal processes and we already know what their valnes are from Eq. (3.1): Q2 = Way = NRT[In(V/Va)], Qi = Wea = NRT; [In (Va/Ve)] = =[n(Ve/Va)]- (Qx is negative.) ‘The efficiency can now be written in terms of the volumes at the different states as par 4 Lillalva/ Ve) Ta[ln(Vi/Vo)] ‘The path from states b to ¢ and from a to d are both adiabatic and reversible. For a reversible adiabatic process wo know that PV7 = constant. Using the ideal gas equation of state, we have TV7~! = constant. Along curve b-c, therefore, T,V,)~! = T,V_'"*. Along the curve d-a, TVg~! = T1Vj~'. Thus, ey (3/1) )" Va _ Va zt A) ( Ve)", which means that 44 = Ye ( (h/t) Vs ve Ve Comparing the expression for thermal efficiency Eq. (3.4) with Bq. (3.5) shows two consequences. First, the heats received and rejected are related to the temperatures of the isothermal parts of the eyele by Qi Qe 5 ete a (3.6) Second, the efficiency of a Carnot cycle is given compactly by qh Carnot eyele efficiency. (7) The efficiency can be 100% only if the temperature at which the heat is rejected is zero, The heat and work transfers to and from the system are shown schematically in Figure 3.5. 34, REFRIGERATORS AND HEAT PUMPS 65 Muddy Points ince 1 = 1—T1/T2, looking at the P-V’ graph, does that mean the farther apart the 71, T; isotherms are, the greater efficiency? And that if they were very close, it would be very inefficient? (MP 3.2) In the Carnot cycle, why are we only dealing with volume changes and not pressure changes on the adiabats and isotherms? (MP 3.3) Is there a physical application for the Carnot cycle? Can we design a Carnot engine for a propulsion device? (MP 3.4) How do we know which cycles to use as models for real processes? (MP 3.5) 3.4 Refrigerators and Heat Pumps ‘The Carnot eycle has heen used for power, but we can also run it in reverse. If so, there is now net work into the system and net heat: out of the system, ‘There will be a quantity of heat Qo rejected at the higher temperature and a quantity of heat Qy absorbed at the lower temperature. ‘The former of these is negative according to our convention and the latter is positive. The result is that work is done on the system, heat is extracted from a low temperature source and rejected to a high temperature source. The words “low” and “high” are relative and the low temperature source might be a crowded classroom on a hot day, with the heat extraction being used to cool the room, The cycle and the heat and work transfers are indicated in Figure 3.6, In this mode of operation the cycle works as a refrigerator or heat pump. “What we pay for” is the work, and “what we get” is the amount of heat extracted. A metric for devices of this type is the coefficient of performance, defined a & Onelficintt. of performanoe Sop ey Figure 3.6: Operation of a Carnot refrigerator For a Carnot eycle we know the ratios of heat in to heat ont when the eyele isn and, since the eyele is reversible, these ratios are the same when the cyele is n m forward reverse. 66 CHAPTER [HE FIRST LAW APPLIED TO ENGINEERING CYCLES ‘The coefficient of performance is thus given in terms of the absolute temperatures as Coefficient of performance = ‘This ean be mneh larger than unit The Carnot cycles that have been drawn are based 0 working media, however, they will look different. We will see an example when we discuss ‘two-phase situations. What is the same whatever the medium is the efficiency for all Carnot cycles operating between the same two temperatures. al gas behavior. For different Refrigerator Hardware pically the thermodynamic system in a refrigerator analysis will be a working fluid, a refrigerant, that circulates around a loop, as shown in Figure 3.7. The internal energy (and temperature) of the refrigerant is alternately raised and lowered by the devices in the loop. The working fluid is colder than the refrigerator air at one point and hotter than the surroundings at another point. ‘Thus heat will flow in the appropriate direction, as shown by the two arrows in the heat exchangers. heeriea! Energy te Figure 3.7: Schematic of a domestic refrigerator Starting in the upper right hand corner of the diagram, we describe the process in more detail. First the refrigerant passes through a small turbine or through an expansion valve. In these devices, work is done by the refrigerant so its internal energy is lowered to a point where the temperature of the refrigerant is lower than that of the air in the refrigerator. A heat exchanger is used to transfer energy from the inside of the refrigerator to the cold refrigerant. This lowers the internal energy of the inside and raises the internal energy of the refrigerant. ‘Then a pimp or compressor is used to do work on the refrigerant, adding additional energy to it and thus further raising its internal energy. Electrical energy is used to drive the pump or compressor. The internal energy of the refrigerant is raised to a poittt 3.5. THE INTERNAL COMBUSTION ENGINE (OTTO CYCLE) or where its temperature is hotter than the temperature of the surroundings. ‘The refrigerant is then passed through a heat exchanger (often coils at the back of the refrigerator) so that energy is transferred from the refrigerant to the surroundings. As a result, the internal energy of the refrigerant is reduced and the internal energy of the surroundings is inereased. It is at this point where the internal energy of the contents of the refrigerator and the energy rroundings. The refrigerant used to drive the compressor or pump are transferred to th then continnes on to the turbine or expansion valve, repeating the cycle. 3.5 The Internal combustion engine (Otto Cycle) [VW, S & B: 9.13] The Otto cycle is a set of processes used by spark ignition internal combustion engines (2-stroke or 4-stroke cycles). These engines a) ingest a mixture of fuel and air, b) compres: it, ¢) cause it to react, thus effectively adding heat through converting chemical energy into thermal energy, d) expand the combustion products, and then e) eject the combustion products and replace them with a new charge of fuel and air. The different processes are shown in Figure 3.8: 1. Intake stroke, gasoline vapor and air drawn into engine (5 — 1). 2. Compression stroke, p, T’ increase (1 + 2). 3. Combustion (spark), short time, essentially constant volume (2+ 3). Model: heat absorbed from a series of reservoirs at temperatures T; to Ts. 4. Power stroke: expansion (3 — 4), 5. Valve exhaust: valve opens, gas escapes. 6. (4+ 1) Model: rejection of heat to series of reservoirs at temperatures Ty to Ty 7. Exhaust stroke, piston pushes remaining combustion products out of chamber (1 + 5). We model the processes as all acting on a fixed mass of air contained in a piston-cylinder arrangement, as shown in Figure 3.10. The actual cycle does not have the sharp transitions between the different processes that the ideal eycle has, and might be as sketched in Figure 3.9. 3.5.1 Efficiency of an ideal Otto cycle ‘The starting point is the general expression for the thermal efficiency of a cycle: _ Qu+Q 4 Qn @ The convention, as previously, is that heat exchange is positive if heat is flowing, into the system or engine, so Qz, is negative. The heat absorbed occurs during combustion when alt 68 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES Vi=% Figure 3.8: The ideal Otto cycle Figure 3.9: Sketch of an actual Otto eyele it volume. ‘The heat absorbed can be related to the the spark occurs, roughly at const temperature change from state 2 to state 3 Qu =Qer= Aas (Wea =0) fs = CodT = Cy(Ts ~ Ta). I ‘The heat rejected is given by (for a perfect gas with constant specific heats) Qt = Qu =AUy = CT - Ts). Substituting the expressions for the heat absorbed and rejected in the expression for thermal efficiency yields Th 3.5, THE INTERNAL COMBUSTION ENGINE (OTTO CYCLE) 69 PRORG Oo © O08 Oo O Figure 3.10: Piston and valves in a four-stroke internal combustion engine We can simplify the above expression using the fact that the processes from 1 to 2 and from 3 to 4 are isentropic: nv TW =nvy" (04—TiVP* = (1s — Tay TT _ (V2) h-hh” \v ‘The quantity Vi/V2 = r is called the compression ratio. In terms of compression ratio, the ‘The ideal Otto cycle efficiency is shown as a function of the compression ratio in Fig- ture 3.11, As the compression ratio, r, increases, yous increases, but so does Ta. If T is too high, the mixture will ignite without a spark (at the wrong location in the cycle). 3.5.2 Engine work, rate of work per unit enthalpy flux ‘The non-dimensional ratio of work cone (the power) to the enthalpy flux through the engine is given by Power Ww Enthalpy flux ~ tieyT) 70 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES $s ‘Compson rio. Figure 3.11: Ideal Otto cycle thermal efficiency ‘There is often a desire to increase this quantity, because it means a smaller engine for the same power. The heat input is given by Qea = tiger brvet where © Alyuot is the heat of reaction, Le. the chemical energy liberated per unit mass of fuel, © ringye is the fuel mass flow rate, ‘The non-dimensional power is Wo _ tira Mio fy tine, th GT sin this equation, evaluated at stoichiometric conditions are: in 8 Aviv 4 10 * 00% 2s" Muddy Points 3.6. DIESEL CYCLE 7 v, Vy W=V; Figure 3.12: ‘The ideal Diesel cycle 3.6 Diesel Cycle The Die into the | cycle is a compression ignition (rather than spark ignition) engine. Fuel is sprayed inder at P (high pressure) when the compression is complete, and there is ignition without a spark, An idealized Diesel engine cycle is shown in Figure 3.12. ‘The thermal efficiency is given by CATs GOs Mien = © h: Os/= ‘This eyele can operate with a higher compression ratio than the Otto eycle because only air is compressed and there is no risk of auto-ignition of the fuel. Although for a given compression ratio the Otto eycle has higher efficiency, because the Diesel engine can be operated to higher compression ratio, the engine can actually have higher efficiency than an Otto eycle when both are operated at compression ratios that might be achieved in practice. Muddy Points When and where do we use cy and cp? Some defini dU = cya? (MP 3.8) Explanation of the above comparison between Diesel and Otto. (MP 3.9) ns use dU = cud’ Is it ever 3.7. Brayton Cycle [VW, $ & B: 9.8-9.9, 9.12] ‘The Brayton cycle (or Joule eycle) reps cycle consists of four processes, as show: seuits the operation of a gas turbine engine. 1 Figure 3.13 alongside a sketch of an engi ¢ a~b Adiabatic, quasi-static (or reversible) compression in the inlet and compressor; © b-c Constant pressure fiiel combustion (idealized as constant pressure heat addition); 2 CHAPTER Figure 3.13: Sketch of the jet engine components and corresponding thermodynamic states © c-d Adiabatic, quasi-static (or reversible) expansion in the turbine and exhaust nozzle, with whieh we 1, take some work out of the air and use it to drive the compressor, and 2. take the remaining work out and use it to accelerate fluid for jet propulsi to turn a generator for electrical power generation; © da Cool the air at constant pressure back to its initial condition. ‘The components of a Brayton cycle device for jet propulsion are shown in Figure 3.14 ‘We will typically represent these components schematically, as in Figure 3.15. In practice, real Brayton eycles take one of two forms. Figure 3.16(a) shows an “open” cycle, where the working fluid enters and then exits the device. ‘This is the way a jet propulsion cycle works. Figure 3.16(b) shows the alternative, a closed cycle, which recirculates the working fluid. Closed cycles are used, for example, in space power generation, Muddy Points ‘Would it be practical to run a Brayton cycle in reverse and use it as refrigerator? (MP 3.10) 3.7.1 Work and Efficiency ‘The objective now is to find the work done, the heat absorbed, and the thermal efficiency of the cycle. ‘Tracing the path shown around the cycle from a-be-d and back to a, the first law gives (writing the equation in terms of a unit mass), Augteta =0= (2 +01 —w Here Au is zero because w is a function of state, and any eyele returns the system to its starting state?. The net work done is therefore w=etay This fact is often useful for solving thermodynamic eyelesin different ways: For instance, in this example ‘we eoitld surn the work terms all around the cycle. Instead we will consider the difference between the heat added to the eyele in process be, and the heat rejected by: the eyele in process dha. 3.7, BRAYTON CYCLE 3 Figure 3.14: Schematics of typical military gas turbine engines. Top: turbojet with after- burning, bottom: GE F404 low bypass ratio turbofan with afterburning (Fill and Peterson, 1992). where q1, g are defined as heat received by the system (q1 is negative). We thus need to evaluate the heat transferred in processes b-e and dea. For a constant pressure, quasi-static process the heat exchange per unit mass is th = cya = dq, oF [ddlcontane P= dhe We can see this by writing the first law in terms of enthalpy’ (see Section 2.3 remembering the definition of ep. The heat exchange can be expressed in te states. Treating the working fluid as a perfect gas with constant specific heats addition from the combustor, ) or by. of enthalpy differences between the relevant for the heat 2 = he — hy = ep(Te— fe). "The heat rejected is, similarly, = ha ~ ha = p(T ~ Ta) "The net work per w it mass is given by Net work per unit mass = q1 + q = ¢p{(Ie— 1h) + (Ta —Ta)]- ‘The thermal efficiency of the Brayton eyele can now be expressed in terms of the tem- peratures: T)) _y_ (la= Ta) _ th) 4 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES 2, Rauivatent heat ransfer at constant pressure 2 3 ‘Compressor FSW. Wet Turbine i 4 G, Equivalent hea transfer 2 at Constant pressure ure 3.15; Thermodynamic model of gas turbine engine cycle for power generation On Puck [Contain Ties camer Tat Air Products To. (8) Open eyete operation (b) Closed eyele operation Figure 3.16: Options for operating Brayton cycle gas turt ne engines ‘To proceed further, we need to examine the relationships between the different temper- atures. We know that points a and d are on a constant pressure process as are points b and ¢, and P, = Py; Py = Pz. The other two legs of the eycle are adiabatic and reversible, so Pa_P, y\VOY 7710-0 fem (Ft -(# z mn > (rz) ~(a) Ty {Ty , or, finally, Ta/7 'e/Tp. Using this relation in the expression ‘y, Eq. (3.8) yields an expression for the thermal efficiency of a Brayton ‘Therefore Ta/Te for thermal e' cycle: — Tumaviae gy Ideal Brayton cycle efficiency: nx 7m 3.7, BRAYTON CYCLE noua Ce es Figure 3.17: Gas turbine engine pressures and temperatures: The temperature ratio across the compressor, T,/T, = TR. In terms of compressor temperature ratio, and using the relation for an adiabatic reversible process we can write the efficiency in terms of the compressor (and eycle) pressure ratio, whieh is the parameter commonly used: l ~ PRO-DA a (3.10) Figure 3.17 shows pressures and temperatures through a gas turbine engine (the PW000, which powers the 747 and the 767). Equation (3.10) says that for a high eycle efficiency, the pressure ratio of the eyele should icreased. This trend is plotted in Figure 3.19. Fignre 3.18 shows the history of aircraft engine pressure ratio versus entry into service, and it can be seen that there has been a lange be increase in cycle pressure ratio. The thermodynamic concepts apply to the behavior of real aerospace devices! Muddy Points When flow is accelerated in a nozzle, doesn’t that reduce the internal energy of the flow and therefore the enthalpy? (MP 3.11) Why do we say the combustion in a gas turbine engine is constant pressure? (MP 3.12) Why is the Brayton cycle less efficient than the Carnot cycle? (MP 3.13) If the gas undergoes constant pressure cooling in the exhaust outside the engine, within the system boundary? (MP 3.14) Does it matter what labels we put on the corners of the cycle or not? (MP 3.15) Is the work done in the compressor always equal to the work done in the turbine plus work cout (for a Brayton cyle)? (MP 3.16) that still 76 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES 40 went 6 @ 5 5 = & 2 é : 8 E 1960 1970 1980 1990 2000 ‘Year of Certification Figure 3.18: Gas turbine engine pressure ratio trends (Janes Aeroengines, 1998) 3.7, BRAYTON CYCLE 7 a Poe bc: oc i er er a onrstr resi ao Figure 3.19: Trend of Brayton cycle thermal efficiency with compressor pressure ratio Figure 3.20: Rolls-Royce high temperature technology 3.7.2. Gas Turbine Technology and Thermodynamics ‘The turbine entry temperature, T., is fixed by materials technology and cost. (If the tem- perature is too high, the blades fail.) Figures 3.20 and 3.21 show the progression of the turbine entry temperatures in aeroengines. Figure 3.20 is from Rolls Royce and Figure 3.21 is from Pratt & Whitney. Note the relation between the gas temperature coming into the turbine blades and the blade melting temperature. For a given level of turbine technology (in other words given maximum temperature) a design question is what should the compressor TR be? What criterion should be used to decide this? Maxinum thermal efficiency? Maximum work? We examine this isstie below. ‘The problem is posed in Figure 3.22, which shows two Brayton eycles. For maximum efficiency we would like TR as high as possible. This means that the compressor exit tem- perature approaches the turbine entry temperature. The net work will be less than the heat received; as T, > T; the heat received approaches zero and so does the net work. The net work in the eyele can also be expressed as [ Pdv, evaluated in traversing the 78 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES lade cooling technology [Pratt & Whitney] Figure 3.22: Efficioney and work of two Brayton cycle engines 3.7. BRAYTON CYCLE 79 cycle. This is the area enclosed by the curves, which is seen to approach zero as Ti, + Te jon from either of these arguments is that a cycle designed for maximum, icy is not very useful in that the work (power) we get out of it is zero. A more useful criterion is that of maximum work per unit mass (maximum power per unit mass flow). ‘This leads to compact propulsion devices. ‘The work per unit mass is given by: Work/unit mass = e,[(Ze—T,) — Tu -Ta)} ‘the maximum turbine i where T. let temperature (a design constraint) and Ty is atmo- spheric temperature. ‘The design variable is the compressor exit temperature, Jj, and to find the maximum as this is varied, we differentiate the expression for work with respect to Tj: Work ame) a, dy an, ~ ? \any at, * ay ‘The first and the fourth terms on the right hand side of the above equation are both zero (the turbine entry temperature is fixed, as is the atmospheric temperature). ‘The maximum work occurs where the derivative of work with respect to Tj, is zero: aWork - a, om) To use Eq, (3.11), we need to relate Zy and Z,. We know that eR Hence, Plugging this expression for the derivative into Eq. (3.11) gives the compressor exit, temperature for maximum work as 1, = yTaT-. In terms of temperature ratio, [Te Compressor temperature ratio for maximum work: 3° = 76 The condition for maximum work in a Brayton cycle is different than that for maximum efficiency. The role of the temperature ratio can be seen if we examine the work per unit mass which is delivered at this condition: oy [t- VER~ Fete + 1]. Ratioing all temperatures to the engine inlet temperature, Work/unit mass Work/unit mass = ¢yTa [i -2y me ol. To find the power the engine can produce, we need to multiply the work per unit mass by the mass flow rate: 80. CHAPTER yele. (3.12) Power = nt (Lhe units anol Watts.) ‘The trend of work output vs. compressor pressure rati TR = Te/Ta, is shown in Figure 3.23 for different temperature ratios Brayton Cycle Work 2s Specie Werke wep os. ° 10 2 30 40 50 ‘Compressor Pressure Ratio Figure 3.23: Trend of cycle work with compressor pressure ratio, for different temperature ratios TR = e/Ta 24 shows the expression for power of an ideal eycle compared with data from actual jet engines. Figure 3.24(a) shows the gas turbine engine layout including the core (compressor, burner, and turbine). Figure 3.24(b) shows the core power for a number of different engines as a function of the turbine rotor entry temperature. The equation in the figure for horsepower (HP) is the same as that which we just derived, except for the Figure 3 3.7, BRAYTON CYCLE 81 i 2 = = =3 i @: me. Figure 3.24: Aeroengine core power [Koff/Meese, 1995 conversion factors. The analysis not only shows the qualitative trend very well but captures much of the quantitative behavior too. A final comment (for this section) on Brayton eycles concerns the value of the thermal efficiency. ‘The Brayton cyele thermal efficiency contains the ratio of the compressor exit ‘temperature to atmospheric temperature, so that the ratio is not based on the highest tem- perature in the cycle, as the Carnot efficiency is. For a given maximum cycle temperature, the Brayton cycle is therefore less efficient than a Camot cycle. Muddy Points What are the units of w in power = sinw? (MP 3.17) Question about the assumptions made in the Brayton cycle for maximum efficiency and maximum work (MP 3.18) You said that for a gas turbine engine modeled as a Brayton cycle the work done is w = qi-+42, where go is the heat added and «1 is the heat rejected. Does this suggest that the work that you et out of the engine doesn’t depend on how good your compressor and turbine are?. ..since the compression and expansion were modeled as adiabatic. (MP 3.19) 3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet A schematic of a ramjet is given in Figure 3.25, In the ramjet there are “no moving part device are: hhe process: s that occur in this propulsion © 0 +3; Isentropic diffusion (slowing down) and compression, with a decrease in Mach number, Mo > Ms < 1. © 3+ 4: Constant pressure combustion. © 4 5: Isentropie expansion through the nozzle. 82 CHAPTER 3. THE FIRST LAW APPLIED TO ENGINEERING CYCLES Figure 3.25: Ieal ramjet [J. L. Kerrebrock, Aéreraft Engines and Gas Turbines] ‘The ramjet thermodynamic cycle efficiency can be written in terms of flight Mach num- ber, Mo, as follows: "Brayton = 1 ~ and Sce also Section 11.6.3 for other figures of merit. Muddy Points Why don't we like the numbers 1 and 2 for the stations? Why do we go 0-3? (MP 3.20) For the Brayton cycle efficiency, why does 7's = Tig? (MP 3.21) 3.7.4 MIT Cogenerator MIT operates a Brayton cycle power generator on campus. For more information, see the website at https://cogen.mit.edu/ctg.cin 3.8 Muddiest points on Chapter 3 MP 3.1 How can we idealize fuel addition as heat addition: The validity of an approximation rests on what the answer is going to be used for. We are seeking basically only one item concerning combustor exit conditions, namely the exit temperature or the exit enthalpy. The final state is independent of how we add the heat, and depends only on whether we add the heat. If it is done from an electrical heater or from combustion, and if we neglect the change in the constitution of the gas due to the combustion products (most of the gas is nitrogen) the enthalpy rise is the same no matter how the temperature rise is achieved. 3.8, MUDDIES ‘T POINTS ON CHAPTER ?? 83 MP 32 Since » = 1 — T/T looking at the P-V graph, does that mean the farther apart the 1}, T2 isotherms are, the greater the efficiency? And that if they were very close, it would be very inefficient? This is correct. However, there is a limit on the maximum achievable efficiency. We cannot convert the absorbed heat into 100% work, thar is, we always must reject some amount of heat. The amount of heat we must reject is Qr= Thus for given values of Ta and Qa. Qre depends only on the temperature of the cold reservoir T,, which is limited by the temperatures naturally available 10 us. These temperatures are all well above absolute zero, and there are no means to reduce Qr to negligible values. The consequence of this is that the Carnot eycle efficiency cannot approach one (11 = 1 only if Qre = 0. which is not possible). QuTi/T2 (see notes for derivation). MP3.3 In the Carnot cycle, why are we only dealing with volume changes and not pressure changes on the adiabats and isotherm We are not neglecting the pressure terms and we are also dealing with pressure changes. On the adiabats we know that dq = 0 (adiabatic process), so that for reversible processes we can write the first law as du = —Pdv and, using enthalpy, also as dh = vdP. With dh = cya and du = cy? {for an ideal gas, we can write the ratio of dh/du as dh[du —(wdP)/(Pdo) lee By arranging terms we obtain aP/P = —ydv/\. Fora process we can integrate from 1 102 and get Pav’ = wv}, or Pu! = const. This relation shows how pressure and volume changes are related to one another during an adiabatic reversible process. During an isothermal process, the temperature stays constant. Using the equation of state for an ideal gas, Pv = RP, we find that Pv = const on an isotherm. Again, this relation tells us how pressure changes are related to volume changes during an isothermal process. Note that in the P-V diagram, adiabats (Pv? = constant) are steeper curves than isotherms (Pv = constant). MP 3.4 Is there a physical application for the Carnot eycle? Can we design a Carnot engine for a propulsion device? We will see that Carnot cycles are the best we can do in terms of efficiency. A constant temper- ature heat transfer process is, however, difficult to attain in practice for devices in which high rates of power are required. The main role of she Carnot engine is therefore as a standard against which all other eycles are compared and which shows us the direction in which design of efficient eycles should go. MP 3.5 How do we know which cycles to use as models for real processes? We have discussed this briefly for the Brayton cycle, in that we looked at the approximation that was made in saying heat addition occurred at constant pressure. You can also see that the Carnot st CHAPTER 3. THE FIR. LAW APPLIED TO ENGINEERING CYCLES cycle is not a good descriptor of a gas turbine engine! We will look further at this general point, not only for the Brayton cycle, but also for the Rankine eycle and for some internal combustion engine cycles. Twill try to make clear what are the approximations and why the cycle under study is being used as a model, MP 3.6 How is Aljueicateulated? For now, we rely on tabulated values. In the lectures accompanying Chapter 15 of the notes, we will see how one can calculate the heat, Ahjueu liberated in a given reaction, MP 3.7 What are “stoichiometric conditions?” Stoichiometric conditions are those in which the proportions of fuel and air are such that there is not an excess of either one in the combustion reaction — all the fuel is burned, and all the air (oxidizer) is used up in the reaction. See Chapter 15 MP3.8 When and where do we use c, and cy? Some definitions use dU = ed. Is it ever dU = cy? The answer is no. The definitions of cy and ¢y are derived in the notes in Section 2.4. cp is the specific heat at constant pressure and for an ideal gas dh = cyAlT always holds. Similarly cy is the specific heat at constant volume and for an ideal gas du = cy always holds. A discussion of this ven in the notes in Section 24 If you think about how you would measure the specific heat | Ljinei ~ Ttict) for a certain known change of state you could do the following experiments. For a process during which heat Aq is transferred (reversibly) and the volume stays constant (e.g. a rigid, closed container filled with a substance, or the heat transfer in an Otto engine during combustion — the piston is near the top-dead-center and the volume is approximately constant for the heat transfer) the first law is du = dq since v = const. Using the definition du = cya we obtain for the specific heat at constant volume eo = Ag/AT, where both the heat transferred Aq and the temperature difference AT’ can be measured. Similarly we can do an experiment involving a process where the pressure is kept constant during the reversible heat transfer Aq (e.g. a rigid container filled with a substance that is closed by a lid with a certain weight, or the heat transfer in a jet engine combustor where the pressure is approximately constant during heat addition). The first law can be written in terms of enthalpy as dh — vdp = dg, and since p = const we obiain dh = dg. Using the definition dh = epi’ we obtain for the specific heat at constant pressure ey = Aa/ AT. MP 3.9 Explanation of the above comparison between Diesel and Otto. Basically we can operate the diesel cycle at much higher compression ratio than the Otto cycle because only air is compressed and we don't run into the auto-ignition problem (knocking problem). Because of the higher compression ratios in the diesel engine we get higher efficiencies. MP 3.10 Would it be practical to run a Brayton cycle in reverse and use it as refrigerator? Yes. In fact people in Cryogenics use reversed Brayton cycles to cool down systems where very ow temperatures are required (¢.g., space applications, liquefaction of propellants). One major 3.8, MUDDIES ‘T POINTS ON CHAPTER ?? 85 difference between a regular Brayton eycle (such as a jet-engine or a gas-turbine) and a reversed Brayton cycle is the working fluid. In order to make a reversed Brayton cycle practical we have to choose a working fluid that is appropriate for the application. Exiremely low temperatures can be achieved when using a regenerator —a heat exchanger that preheats the fluid before it enters the compressor and cools the fluid further down before it enters the turbine. In this configuration the fluid is expanded to much lower temperatures, and more heat can be absorbed from the cooling compartment. MP 3.11 When flow is accelerated in a nozzle, doesn't that reduce the internal energy of the flow and therefore the enthalpy? Indeed both enthalpy and internal energy are reduced. The stagnation enthalpy is the quantity that is constant. MP 3.12 Why do we say that the combustion in a gas turbine engine iy at constant pressure? This is an approximation, and a key question is indeed how accurate itis and what the justifica- tion is. The pressure change in the combustor can be analyzed using the one-dimensional compress- ible flow equations. The momentum equation is dP = —pede, where c is the velocity. If we divide both sides by P, we obvain: aP_ 12 de__yeede P >P]pe where a is the speed of sound and M is the Mach number. Changes in velocity are due to changes in density and in flow-through area A, as given by the one-dimensional continuity equation de ao = mre, @ ¢ ee pea = constant. Hence Inp+Inc+In A = constant. Differentiating, Velocity changes are therefore related to area changes (geometry) and density changes (basically heat input). For a gas turbine combustion process the change in density is comparable with (a Significant fraction of) the initial density and the area change is several times the initial area. This ‘means that the change in velocity divided by the initial velocity is roughly of the order of magnitude of unity. The momentum equation thus tells us that for small Mach number (say 0.1) the ratio dP/ P will be much less than one, so that the pressure can be approximated as constant. In reality the pressure does drop in the combustor, but the overall drop from inlet to exit is about 3-4%, small compared to the initial level of pressure, $o that the approximation of constant pressure is a useful The rapidity of the combustion process does not really have anything to do with this approxima- tion. We could have a process, such as a nozcle, in which there was combustion at the same time that the pressure was dropping. As seen from the momentum equation, the heat addition does not “directly” affect the pressure — changes in pressure are associated with changes in velocity: 86 CHAPTER 3. TO ENGINEERING CY¢ Brayton eyele (solid line) Figure 3.26: Brayton cycle considered as a number of elementary Carnot ¢) the same pressure ratio and therefore the same temperature ratio, which overall cycle temperature ratio, Tiyex/Tain les, all having lower than the MP 3.13 Why is the Brayton cycle less efficient than the Carnot eycle? Consider the Brayton cycle and the corresponding work done as being approximated by a num- ber of elementary Carnot eycles, as shown by the dashed lines in Figure 3.26. All of these Carnot ceyeles have the same pressure ratio, thus the same temperature ratio, and thus the same efficiency. The temperature ratio that figures into the efficiency of the elementary Carnot cycles is the inlet temperature divided by the compressor exit temperature, not the maximum cycle temperature, which is at the combustor exit. The basic reason for the lower efficiency is that heat is absorbed at an average temperature that is lower than the maximum temperature and rejected at an average tem- perature higher than the minimum temperature. We will come back to this important poirt (which ‘has implications for all cycles), but if you cannot wait, see Chapter 6 of the notes. MP 3.14 If the gas undergoes constant pressure cooling in the exhaust outside the engine, is that Still within the system boundary? When we analyze the state changes as we trace them around the cycle, we are viewing the changes in a system, amass of fied identity. Thus we follow the mass as it moves through the device and the cooling of the gas outside the engine is happening to our system. MP 3.15 Does it matter what labels we put on the corners of the cycle or not? Itdoes not matter what labels we use on the comers of the cycle. A eyele is.a series of processes. Independent of where you startin the eycle, it always brings you back to the state where you started. MP 3.16 Js the work done in the compressor abvays equal to the work done in the turbine plus work out (for a Brayton eyle)? NO. The work done in the compressor plus net work out equals the total turbine work. Using the 1 aw, the net work we get out of the Brayton cycle is w ‘pl(Ta~ Ta) + (Le ~ To) (see notes for details). Rearranging the temperatures we can also write mean 3.8, MUDDIES ‘T POINTS ON CHAPTER ?? 87. fa) — (Te — Thus the net work is the difference between the enthalpy drop across the turbine (we get work from the turbine) and the enthalpy rise through the compressor (we have 0 supply work 10 the compressor). w= el(Te— fa)] = Arcurbine — Abeompressor MP 3.17 What are the units of win power = rinw? The units of power are J/s (ki/s, MU/s) or Watis (KW, MW). The mass flow is kg/s. The units of 1w, work per unit mass, are thus J/kg. For the aeroengine, we can think of a given diameter (frontal area) as implying a given mass flow (think of a given Mach number and hence a given ratio of flow 10 choked flow). If so, for a given fan diameter power scales directly as work per unit mass. MP 3.18 Question about the assumptions made in the Brayton cycle for maximum efficiency and ‘maximum work. We have first derived a general expression for the thermal efficiency of an ideal Brayton cycle (see Equation 3.10), The assumptions we made for the cycle were that both the compressor and turbine are ideal, such that they can be modeled adiabatic and reversible. We then looked at possi- ble ideal Brayton eycles that would yield (A) maximum efficiency and (B) maximum work, keeping the assumptions of an ideal eycle (the assumptions of adiabatic and reversible compression and expansion stem from the choice of an ideal cycle). One way to construct an ideal Brayton cycle in the P-V diagram is to choose the inlet temper- ature Ta and inlet pressure Pa, the compressor pressure ratio P)/Pa or temperature ratio To/Tax and the turbine inlet temperature Ts. Apart from setting the inlet conditions (these mainly depend (on the flight altitude and Mach number and the day), we decided to,fx the turbine inlet temperature (fixed by material technology or cost). So the only two “floating” cycle parameters that remain to be defined are the compressor exit temperature Ty and the turbine exit temperature Ty. Looking «at Equation 3.10 we know that the higher the compressor temperature ratio Ts/Tq the higher the thermal efficiency. So, for (A) maximum efficiency we would choose the compressor exit temperature as high as possible, that is in the limit Tj, = T.. Constructing this cycle in the P-V diagram and letting Ty approach Te shows that the area enclosed by the cycle, or in other words the net work, becomes zero. Thus a cycle constructed for maximum efficiency, under the given inlet conditions and constraints on T., is not very useful because we don’t get any work out of it. For the derivation of Ty for maximum work (keeping T fixed as above), see notes for details. MP 3.19 You said that for a gas turbine engine modeled as a Brayton cycle the work done is w = qi +m where qo is the heat added and q; is the heat rejected. Does this suggest that the work that you get out of the engine doesn’t depend on how good your compressor and turbine are?...since the compression and expansion were modeled as adiabatic. ng the 1 law, the net work we get out of the cycle is w= a +a = epl(Ta— (see notes for details). Rearranging the temperatures we can also write uw) + (Ze ~ Th) w = ep[(Te = Ta) = (To ~ Ta)] = Abreurvine ~ Abtcompressor: 88 CHAPTER 3. THE FIR. LAW APPLIED TO ENGINEERING CYCLES Thus the net work is the difference between the enthalpy drop across the turbine (we get work from the turbine) and the enthalpy rise through the compressor (we have 10 supply work to the ‘compressor, this is done through the drive shaft that connects turbine and compressor). In class we analyzed an ideal Brayton cycle with the assumptions of adiabatic reversible com- pression and expansion processes, meaning that the work done by the turbine is the maximum work we can get from the given turbine (operating between T- and Ty), and the work needed to drive the given compressor is the minimum work required. In the assumptions the emphasis is put on reversible rather than adiabatic. For real engines the assumption of adiabatic flow through the com- pressor and turbine still holds. This is an approximation — the surface inside the compressor or turbine where heat can be transferred is much smaller than the mass flow of the fluid moving through the machine so that the heat transfer is negligible — we will discuss the different concepts of heat transfer later in class. However the compression and expansion processes in real engines are ire- versible due 10 non-ideal behavior and loss mechanisms occurring in the turbomachinery flow. Thus the thermal efficiency and work for a real jet engine with losses depend on the component efficiencies of turbine and compressor and are less than for an ideal jet engine. We will discuss these component efficiencies in more detail in class. MP 3.20 Why don’t we like the numbers 1 and 2 for the stations? Why do we go 0-3? A common convention in the industry is that station 0 is far upstream, station 1 is after the shock in the inlet (if there is one), station 2 is at inlet to the compressor (after the inletdiffuser) and station 3 is after the compressor. In class, when we examined the ramjet we considered no changes in stagnation pressure between 0 and 2, so have used 0 as the initial state for the compression process. It would be more precise to differentiate between stations 0 and 2, and 1 will do this where ‘appropriate. MP 3.21 For the Brayton cycle efficiency, why does Ts = Tyo? The ramjet is operating as a Brayton eycle where mp = 1—Tinice/Tcompressr este For the ramjet discussed in class the inlet temperature is To and since there is no compressor (no moving parts) the only compression we get is from diffusion. We assumed isentropic diffusion in the diffuser and found {for very low Mach numbers that the diffuser exit or combustor inlet temperature Ts is Tis. From the first law we know that for a steady, adiabatic flow where no work is done the stagnation enthalpy ‘Stays constant. Assuming a perfect gas with constant specific heats we thus get Ty = Tis = Ts. So we can write for the ramjet thermal efficiency me Part II THE SECOND LAW OF THERMODYNAMICS Chapter 4 Background to the Second Law of Thermodynamics [VN Chapters 2, 3, 4] So far we have dealt largely with ideal situations involving qua: frictionless pistons). We will now consider more general situations. i-statie processes (i.e. 4.1 Reversibility and Irreversibility in Natural Processes ality” in nature. We start by examining a flywheel in a fluid filled insulated enclosure as shown in Figure 4.1 A question to be asked is whether we could start with state B and then let events proceed to state A? Why or why not? ‘The first law does not prohibit this. ‘The characteristics of state A are that the energy is in an organized form, the molecules in the flywheel have some circular motion, and we could extract some work by using the flywheel kinetic energy to lift a weight. In state B, in contrast, the energy is associated with disorganized motion on a molecular scale. The temperature of the fluid and flywheel are higher than in state A, so we could probably get some work ont by using a Carnot cycle, but it would be much less than the work we could extract in state A. There is a qualitative We wish to characterize the “direction” of natural processes; there is a basic “directic difference between these states, which we need to be able to describe more precisely. x & hs 2 iit < State A: flywheel spinning, Sate Be Ayes stationary system coo! Figure 4.1: Flywheel in insulated enclosure at initial and final states 92 CHAPTER 4. BACKGROUND TO THE SECOND LAW OF THERMODYNAMICS Muddy Points Why is the al ity to do work decreased in B? How do we know? (MP 4.1) Another example is a system composed of many bricks, half at a high temperature Ty and half at a low temperature 7;,, as shown in Figure 4.2. With the bricks separated thermally, ‘we have the ability to obtain work by running a cycle between the two temperatures. Suppose wwe put two bricks together. Using the first law we can write CT + CT, = 2CTy, (Ta +71)/2 = Tu, where C is the “heat capacity” = AQ/AT. (For solids the heat capacities (specific heats) at constant pressure and constant volume are essentially the same.) We have lost the ability to got work out of these two bricl balfat Ty, hot (igh) halfatT;, low Figure 4.2: Bricks separated by a temperature difference Can we restore the system to the original state without contact with the outside? The answer is no. Can we restore the system to the original state with contact with the ou The answer is yes. We could run a refrigerator to take heat out of one brick and put it into the other, but we would have 10 do work. ‘We can think of the overall process involving the system (the two bricks in an insulated setting) and the surroundings (the rest of the universe) as: # System is changed, © Surroundings are unchanged. ‘The composite system (system and the surroundings) is changed by putting the bricks together. The process is not reversible — there is no way to undo the change and leave no mark on the surroundings. What is the measure of change in the surroundings? 1. Energy? ‘This is conserved. Ability to do work? This is decreased. ‘The measurement and characterization of this type of change ~ of losing the ability to do work ~ is the subject of the second law of thermodynamics. [VW, $ & B: 6.3-6.4] DIFFERENCE BETWEEN F EE AND ISOTHERMAL EXPANS 01 iS 93 4.2 Difference between Free Expansion of a Gas and Re- versible Isothermal Expansion ‘The difference between reversible and irreversible processes is brought out through exam ination of the isothermal expansion of an ideal gas. ‘The question to be asked is what is the difference between the “free expansion” of a gas and the isothermal expansion against a piston? To answer this, we address the steps that we would have to take to reverse, in other words, to undo the process. By free expan on, we mean the unrestrained expansion of a gas into a volume as shown in Figure 4.3. (The restrained expansion is shown in Figure 4.4.) Tuitially all the gas is in the volume designated as Vi with the rest of the insulated enclosure a vacuum. ‘T e total volume (Vi plus the evacuated volume) is V2. At a given time a hole is opened in the partition and the gas rushes through to fill the rest of the enclosure. te 4.3: Froe expansion Figure 4.4: Expansion against a piston During the expansion there is no work exchanged with the surroundings because there is no motion of the boundaries', The enclosure is insulated so there is no heat exchange. ‘The first law tells us therefore that the internal energy is constant (AU = 0). For an ideal gas, the internal energy is a function of temperature only so that the temperature of the gas before the free expansion and after the expansion has been completed is the same, Characterizing the before and after state Tindeed, some ability to do work was lost, for we could have put a piston in the volume and allowed the expansion Of the gosto do work to fuise a weight, 94 CHAPTER 4. BACK ROUND TO THE SECOND LAW 0 THERMODYNAMICS © Before: State 1, V =, T=T) ° Ant State 2, V = Va, 7 Q=W =, so there is no change in the surroundings. ‘Yo restore the original state, i.e, to go back to the original volume at the same temper- ature (Vp + Vj at constant T = 7;) we can compress the gas isothermally (using work from an external agency). We can do this in a quasi-equ as in Figure 4.5, If so the work that we need to do is W = ff PdV. We have evaluated the work in a reversible isothermal expansion (Eq. 3.1), and we can apply the arguments to the case of a reversible isothermal compression. The work done on the system to go from state “2? to state “I” is librium manner, with Paystem © Pexteraals Work done on system = NRT, In (#) : vi f Proatem ff Peter “Pays + dp Figure 4.5: Returning the free expansion to its initial condition From the first law, this amount of heat must also be rejected from the gas to the surround- ings if the temperature of the gas is to remain constant. A schematic of the compression Figure 4.6 in terms of heat and work exchanged is shown process, Wwe) 2 eatou) Figure 4.6; Work and heat exchange in the reversible isothermal compression process At the end of the combined process (free expansion plus reversible compressiot 1. The system has been returned to its initial state (no change in system state). 4.2, DIFFERENCE BETWEEN FREE AND ISOTHERMAL EXPANSIONS 95 4 Block Weight OO Figure 4.7: 100% conversion of work into heat I 2 2 Ge — a Figure 4.8: Work and heat transfer in reversible isothermal expansion 2. The surroundings (us!) did work on the system of magnitude? W. 3. The surroundings received an amount of heat, Q, which is equal to W. 4, The sum of all of these events is that we have converted an amount of work, WW, into an amount of heat, Q, with Wand Q numerically equal in Joules. ‘The net effect is the same as if we let a weight fall and pull a block along a rough surface, as in Figure 4.7. There is 100% conversion of work into heat. ‘The results of the free expansion can be contrasted against a process of isothermal ex- pansion against a pressure dP which is slightly different than that of the system, as shown in Figure 4.8. During the expansion, work is done on the surroundings of magnitude W = f PaV, where P can be taken as the system pressure. As evaluated in Eq. (3.1), the magnitude of the work done by the system is W = NR; In(V/Vi). At the end of the isothermal expansion, therefore 1. The surroundings have received work W, 2. The surroundings have given up heat, Q, numerically equal to W. fote that without our external work, the gas would not move itself back into Vi, even though such a the First Law. spontancous reversal would be compatib 96 CHAPTER 4. BACKGROUND TO THE SECOND LAW OF THERMODYNAMICS free expansion. In fact, because we are doing a transition between the same states along the same path, the work and heat exchange are th examined just above. The overall result when we have restored the system to the quite different for the reversible expansion than for the free expansion, For the reversible expansion, the work we need to do on the system to compress it has the same magnitude as the work we received during the expansion process. Indeed, we could raise a weight during the expansion and then allow it to be lowered during the compression process. Similarly ‘the heat put into the system by us (the surroundings) during the expansion process has the sane magnitude as the heat received by us during the compression process. The result is that when the system has heen restored back to its initial state, so have the surroundings. ‘Uhere is no trace of the overall process on either the system or th meaning of the word “reversible, ‘ame as those for the compression process ial state, however, is wrroundings. ‘That is another Muddy Points With the isothermal reversible expansion is Payternal constant? If so, how can we have Paystem * Pesternat? (MP 4.2) ‘Why is the work done equal to zero in the free expansion? (MP 4.3) Is irreversibility defined by whether or not a mark is left on the outside environment? (MP 4.4) 4.3. Features of reversible processes [VW, S & B: 6.3.6.4] Reversible processes are idealizations or models of real processes. One familiar and widely used example is Bernoulli's equation, which you saw in Unified. They are extremely nseful for defining limits to system or device behavior, for enabling identification of areas in which inefficiencies occur, and in giving targets for design An important feature of a reversible process is that, depending ou the process, it repre- sents the maximum work that can be extracted in going from one state to another, or the minimum work that is needed to create the state change. Let us consider processes that do work, so that we can show that the reversible one produces the maximum work of all possible processes between two states. Por example, suppose we have a thermally insulated cylinder that holds an ideal gas, Figure 4.9. The gas is contained by a thermally insulated massless piston with a stack of many small weights on top of it. Initially the system is in mechanical and thermal equilibrium, Consider the following three processes, shown in Figure 4.10: 1. All of the weights are removed from the piston instantaneously and the gas expands until its volume is increased by a factor of four (a free expansion). 2. Half of the weight is removed from the piston instantaneously, the system is allowed to double in volume, and then the remaining half of the weight is instantaneously removed from the piston and the gas is allowed to expand until its volume is again doubled. . FEATURES OF REVERSIBLE PROCESSES 97 tas Figure 4.9: A piston with weights on top 3. Each small weight is removed from the piston one at a time, so that the pressure inside the cylinder is always in equilibrium with the weight on top of the piston. When the last weight is removed, the volume has increased by a factor of four. ena ret tan Figure 4.10: Getting the most work out of a system requires that the work be extracted reversibly Maximum work (proportional to the area under these curves) is obtained for the quasi- static expansion, *Note that there is a direct inverse relationship between the amount of work received from a process and the degree of irreversibility. To reiterate: # The work done by a system during a reversible process is the maxitnum work we can © The work done on a-system in a re do to achieve that state change. ible process is the minimum work we need to A process must be quasi-static (quasi-equilibrium) to be reversible. ‘This means that the following effects must be absent or negligible: 1. Friction: If Paxernat # Paystem We would have to do net work to bring the system from one volume to another and return it to the initial condition (recall Section 1.3.3.) 2, Free (unrestrained) expansion, 3. Heat transfer through a finite temperature difference, 98 CHAPTER 4. BACKGROUND TO THE SECOND LAW OF THERMODYNAMICS @ Figure 4.11: Heat transfer across a e temperature difference Suppose we have heat transfer from a high temperature to a lower temperature as shown in Figure 4.11, How do we restore the situation to the initial eonditions One thought would be to run a Carnot refrigerator to get an amount of heat, Q, from the lower temperature reservoir to the higher temperature reservoir. We could do this but the surroundings, again us, would need to provide some amount of work (which we could find using our analysis of the Carnot refrigerator). ‘The net (and only) result at the end of the combined process would be a conversion of an amount of work into heat. For reversible heat transfer from a heat reservoir to a system, the temperatures of the system and the reservoir must be Teystem + AT. In other words the difference between the temperatures of the involved in the heat transfer process ean only differ by an infinitesimal amount, Theor rservo ‘two entiti ar. While all natural processes are irreversible to some extent, it cannot be emphasized too strongly that there are a number of engineering situations where the effect of irreversibility can be neglected and the reversible process furnishes an excellewt approximation to reality. The second law, which statement concerning the the next topic we address, allows rreversibility of a gi us to make a quantitative n physical process. Figure 4.12: Nicolas Sadi Carnot (1796-1832), an engineer and an officer in the French army. Camot’s work is all the more remarkable because it was made without benefit of the first law, which was not discovered until 30 years later. [Atkins, The Second Lavi] 44, MUDDIES ‘T POINTS ON CHAPTER ?? 99, Muddy Points Is heat transfer across a finite temperature difference only irreversible if no device is present between the two to harvest the potential difference? (MP 4.5) 4.4 Muddiest Points on Chapter 4 MP 4.1 Why is the ability to do work decreased in B? How do we know? In state A, the energy is in organized form and the molecules move along circular paths around the spinning flywheel. We could get work out this system by using all of the kinetic energy of the flywheel and for example lift a weight with it. The energy of the system in state B (flywheel not spinning) is associated with disorganized motion (on the molecular scale). The temperature in state Bis higher than in state A. We could also extract work from state B by running for example an ideal Carnot cycle between Tp and some heat reservoir at lower temperature. However the work we would get from this ideal Carnot cycle is less than the work we get from state A (all of the kinetic energy), because we must reject some heat when we convert heat into work (we cannot convert heat into 100% work). Although the energy of the system in state A is the same as in state B (we know this from 1" law) the “organization” of the energy is different, and thus the ability to do work is different. MP 4.2 With the isothermal reversible expansion, is Pexternai constant? If so, how can we have Paystem ~ Pesternal? For a reversible process, if the external pressure were constant, there would need to be a force that pushed on the piston so the process could be considered quasi-equilibrium. This force could be us, it could be a system of weights, or it could be any other work receiver. Under these conditions the system pressure would not necessarily be near the external pressure but we would have Paystem © Peaternat Fwork receiver! Apiston. We can of course think of a situation in which the external pressure was varied so it was always close to the system pressure, but that is not necessary. MP 4.3 Why is the work done equal to zero in the free expansion? In this problem, the system is everything inside the rigid container. There is no change in volume, no “dV,” so no work done on the surroundings. Pieces of the gas might be expanding, pushing ‘on other parts of the gas, and doing work locally inside the container (and other pieces might be compressed and thus receive work) during the free expansion process, but we are considering the system asa whole, and there is no net work done. MP 4.4 Is irreversibility defined by whether or not a mark is left on the outside environment? A process is irreversible when there is no way to undo the change without leaving a mark on the surroundings or “the rest of the universe.” In the example with the bricks, we could undo the change by putting a Carnot refrigerator between the bricks (both at Tis after putting them together) and cooling one brick down to T;, and heating the other brick to Ty to restore the initial state. To do this we have to supply work to the refrigerator and we will also reject some heat to the surroundings. Thus we leave a mark on the environment and the process is irreversible. MP 4,5 Is heat transfer across a finite temperature difference only irreversible if no device is present hetween the two to harvest the potential difference? If we have two heat reservoirs at different temperatures, the irreversibility associated with the transfer of heat from one 10 the other is indeed dependent on what is between them. If there is a 100 CHAPTER 4. BACKGROUND TO THE SECOND LAW OF THERMODYNAMICS copper bar between them, all the heat that comes out of the high temperature reservoir goes into the low temperature reservoir, with the result given in Section 5.5 If there were a Carnot cycle between them, some (not all) heat from the high temperature reservoir would be passed on to the Iow temperature reservoir, the process would be reversible, and work would be done. The extent 10 which the process is irreversible for any device can be assessed by computing the total entropy change (device plus surroundings) associated with the heat transfer. Chapter 5 The Second Law of Thermodynamics [VN Chapter 5; VWB&S-6.3, 6.4, Chapter 7] 5.1 Concept and Statements of the Second Law (Why do we need a second law?) ‘The nnresti ar ined expansion, or the temperature equilibration of the two bricks, are fat processes. Suppose you are asked whether you have ever seen the reverse of these processes take place? Do two bricks at a medium temperature ever go to a state where one is hot and one is cold? Will the gas in the unrestrained expansion ever spontaneously return to occupying only the left side of the volume? Experience hints that the answer is no. However, both these processes, unfamiliar though they may be, are compatible with the first law. In other words the first law does not prohibit their occurrence, There thus must be some other “great principle” that describes the direction of natural processes, that tells us which first Jaw compatible processes will not be observed. ‘This is contained in the second law. Like the first law, it is a generalization from an enormous amount of observation. There are several ways in which the second law of thermodynamics can be stated. Listed below are three that are often encountered. As described in class (and as derived in al- most every thermodynamics textbook), although the three may not appear to have much connection with each other, they are equivalent. 1. No process is possible whose sole result is the absorption of heat from a reservoir and the conversion of this heat into work. [Kelvin-Planck statement of the second law] 2. No process is possible whose sole result is the transfer of heat from a cooler to a hotter body. [Clansius statement of the second law] 3. There exists for every system in equilibrium a property called entropy, S, which is a thermodynamic property of a system. For a rey le process, changes in tl are given by dS = (dQreversivie)/T- 101 102 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS Figure 5.1: This is not possible (Kelvin-Planck) 2 ' i | nto 1 1 ~H- A (7, 1 Q Figure 5.2: For Ty < Tz, this is not possible (Clausius) 5.1. CONCEPT AND STATEMENTS OF THE SECOND LAW 103 ‘The entropy change of any system and its surroundings, considered together, is positive and approaches zero for any process which approaches reversibilit AStorai 2 0. For an isolated system, i.e., a system that has no interaction with the surroundings, changes in the system have no effect on the surroundings. In this case, we need to consider the system only, and the first and second laws become: AE stem = 0, ASwysiem > 0. olated system the total energy (£ = U + Kinetic Energy + Potential Energy -+ -..) is constant. The entropy can only increase or, in the limit of a reversible process, remain constant. The limit, Syoeq1 = const or AS. =O, represents the best that can be done. In ther- modynamies, propulsion, and power generation systems we often compare performance to this limit to measure how close to ideal a given process is Alll of these statements are equivalent, but 3 gives a direct, quantitative measure of the departure from reversibility. Entropy is not a familiar concept and it may be helpful to provide some additional rationale for its appearance. If we look at the first law, av = dQ - aw, the term on the left is a funetion of state, while the two terms on the simple compressible substance, however, we can write the work done in a reversible process as dW = PdV , so that aU = dQ — PaV First law for a simple compressible substance, reversible process. ‘Two out of the three terms in this equation are expressed in terms of state variables. Tt seems plausible that we ought to be able to express the third term using state variables as well, but what are the appropriate variables? If so, the term dQ = (){] should perhaps be viewed as analogous to dV = PdV where the parentheses denote an intensive state variable and the square brackets denote an extensive state variable. ‘The second law tells us that the intensive variable is the temperature, T, and the extensive state variable is the entropy, S. ‘The first law for a simple compressible substance in terms of state variables is thus dU = Tds ~ Pav. (6.1) Because Eq. 5.1 includes the second law, it is referred to as the combined first and second law. Because it is written in terms of state variables, it is true for all processes, not just reversible ones. We summarize below some attributes of entropy: 104 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS 1. Entropy is a function of the state of the system and ean be found if any two properties of the system are known, e.g. += (p,T) or 8 = s(T,v) or 8 = s(p, 0). 2. S is an extensive variable. The entropy per unit mass, or specific entropy, is. 3, The units of entropy are Joules per degree Kel are J/K-kg. in (J/K). The units for specific entropy ‘stem and the heat is 4, For a system, dS = dQrev/T, where the numerator is the heat given to th the denominator is the temperature of the system at the location whe received. dS =0 for pure work transfer. Muddy Points Why is dU = TdS — PaV always true? (MP 5.1) What makes Qe. different than dQ? (MP 5.2) 5.2 Axiomatic Statements of the Laws of Thermodynamics 5.2.1 Introduction As a further aid in familiarization with the second law of thermodynamics and the idea of entropy, we draw an analogy with statements made previously concerning quantities that are closer to experience. In particular, we wish to prosent once more the Zeroth and First Laws of thermodynamics and use the same framework for the Second Law. In this so-called “axiomatic formulation,” the parallels between the Zeroth, First and Second Laws will be made explicit. 5.2.2 Zeroth Law Section 1.3.2 presented this observation: Zeroth Law: There exists for every thermodynamic system in equilibrium a prop- erty called temperature. Equality of temperature is a necessary and sufficient condition for thermal equilibrium. The Zeroth law thus defines a property (temperature) and describes its behavior. 5.2.3. First Law Observations also show that for any system there is a property called the energy. The First Law asserts that one must associate such @ property with e ‘vem, "rom notes of Professor F. F.C. Click, California Institute of Technology’ (with minor changes). 5.2. AXIOMATIC STATEME) TS OF THE LAWS OF THERMODYNAMICS 105 First Law: There exists for every thermodynamic system a property called the energy. ‘The change of energy of a system is equal to the mechanical work done on the system in an adiabatic process. In a non-adiabatie process, the change in vray is equal to the heat added to the system minus the mechanical work do: by the system. On the basis of experimental results, therefore, one is led to assert the existence of, two new properties, the temperature and internal energy, which do not arise in ordinary mechanics. Ina similar way, a further remarkable relationship between heat and temperature will be established, and a new property, the entropy, defined. Although this is a much less familiar property, it is to be stressed that the general approach is qnite like that used to establish the Zeroth and First Laws. A general principle and a property associated with any system are extracted from experimental results. Viewed in this way, the entropy should appear 0 more mystical than the internal energy. The increase of entropy in a naturally oceurring process is no less real than the conservation of energy. 5.2.4 Second Law Although all natural processes must take place in accordance with the First Law, the princi- ple of conservation of energy is, by itself, inadequate for an unambiguous description of the behavior of a system. Specifically, there is no mention of the familiar observation that every natural process has in some sense a preferred direction of action. For example, the flow of heat occurs naturally from hotter to colder bodies, in the absence of other influences, bu the reverse flow certainly is not in violation of the First Law. So far as that law is concerned, the initial and final states are symmetrical in a very important respect. The Second Law is essentially different from the First Law; the two prineiples are inde- pendent and cannot in any sense be deduced from one another. ‘Thus, the concept of energy is not sufficient, and a new property must appear. This property can be developed, and the Second Law introduced, in much the same way as the Zeroth and First Laws were presented, By examination of certain observational results, one attempts to extract from experience a law which is supposed to be general; it is elevated to the position of a fundamental axiom to be proved or disproved by subsequent experiments. Within the structure of classical ther- modynamies, there is no proof more fundamental than observations. A statement which can be adopted as the Second Law of thermodynamics is: Second Law: There exists for every thermodynamic system in equilibrium an extensive scalar property called the entropy, S, stich that in an infinitesimal reversible change of state of the system, dS = dQ/T, where 7’ is the absolute temperature and dQ is the amount of heat received by the system. The entropy of a thermally insulated system cannot decrease and is constant if and only if all processes are reversible, As with the Zeroth and First Laws, the existence of a new property is asserted and its behavior is described 106 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS 5.2.5 Reversible Processes In the course of this development, the idea of a completely reversible process is central, and ‘we can recall the definition, “a process is called completely reversible if, after the process has occurred, both the system and its surroundings can be wholly restored by any means to their respective initial states” (first introduced in Section 1.3.3). Especially, it is to be noted that the definition does not, in this form, specify that the reverse path must be identical with the forward path. If the initial states can be restored by any means whatever, the process is by definition completely reversible. If the paths are identical, then one usually calls the process (of the system) reversible, or one may say that the state of the system follows a reversible path. In this path (between two equilibrium states 1 and 2), (i) the system passes through the path followed by the equilibrium states only, and (ii) the system will take the reversed path 2 to 1 by a simple reversal of the work done and heat added, Reversible processes are idealizations not actually encountered. However, they are clearly useful idealizations. For a process to be completely reversible, it is necessary that it be quasi- static and that there be no dissipative influences such as friction and diffusion. The precise (necessary and sufficient) condition to be satisfied if a process is to be reversible is the second part of the Second Law. The criterion as to whether a process is completely reversible must be based on the initial and final states. In the form presented above, the Second Law furnishes a relation between the properties defining the two states, and thereby shows whether a natural process connecting the states is possible. Muddy Points What happens when all the energy in the universe is uniformly spread, i.e., entropy at a maximum? (MP 5.3) 5.3 Combined First and Second Law Expressions The first law, written in a form that is always true: dU = dQ - aw. For reversible processes only, work or heat may be rewritten as aw = Pav, dQ = Tas. Substitution leads to other forms of the first law srue for reversible processes only dU = dQ — PaV,, substituted for a reversible dW dU = TaS ~ dW, substituted for a reversible dQ. (If the substance has other work modes, e.g., stress, str dU = dQ PdV ~ XaY, 5.4, ENTROPY CHANGES IN AN IDEAL GAS 107 where X is a pressure-like quantity, and Y is a volume-like quantity.) Substituting for both dW and dQ in terms of state variables, a rdS—PdV Always true. ‘The above is always true because it is a relation between properties and is now independent of process. In terms of specific quantities: du =Tds— Pdv Combined first and second law (a) or Gibbs equation (a). The combined first and second law expressions are often more usefully written in terms of the enthalpy, or specific enthalpy, h = u+ Po, dh = du + Pdv + vdP Tas — Pdv + Pdv + vdP, using the first law. rds + vdP. ah Or, since v = 1/p, ‘ds + 2 ‘Combined first and second law (b) or Gibbs equation (b)- dh In terms of enthalpy (rather than specific enthalpy) the relation is dH = Td8 + VaP. 5.4 Entropy Changes in an Ideal Gas [VW, $ & B: 6.5- 6.6, 7.1] Many aerospace applications involve flow of gases (e.g., air) and we thus examine the entropy relations for ideal gas behavior. ‘The starting point is form (a) of the combined first and second law du = Tds ~ Pav. For an ideal gas, du = ed. Thus Tas = ed + Pdv or Using the equation of state for an ideal gas (Pv = RT), we can write the entropy change as an expression with only exact differentials: (62) 108 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS We can think of Equation (5.2) as relating the fractional change in temperature to the fractional change of volume, with scale factors cy and AR; if the volume increases without a proportionate decrease in temperature (as in the case of an adiabatic free expansion), then 5 increases. Integrating Equation (5.2) between two states “I” and “2": n ede as=a—a= [ og eR ae hn ne For a perfect gas with constant specific heats As= 82-5 =¢1n (2) +Rln (2). In non-dimensional form (using R/ey = (7 ~ 1)) ce qh 1 Equation 5.3 is in terms of specific quantiti As (2 Ge “AT, ‘This expression gives entropy change in terms of temperature and volume, We can develop an alternative form in terms of pressure and volume, which allows us to examine an assumption we have used. The ideal gas equation of state can be written as As Ts % =n (?) +(y-1)hn (2) ; Entropy change of a perfect gas. (5.3) For N moles of gas, 2 InP+Inv=InR4+ nT. ‘Taking differentials of both sides yields ng the above equation in Eq, (5.2), and making use of the relations cp we find + Rs ep/ew or @ Pe Integrating between two states 1 and 2 -a(Re—(@)-mfR Go Using both sides of (5.4) as exponents we obtain Paul = [Purp = ever, (5.5) CALCULATION OF ENTROPY CHANGE IN SOME BASIC PROCESSES 109 Ta t Ou On Pigure 5.3: Heat transfer from/to a heat reservoir Equation (5.5) describes a general process. For the specific situation in which As = 0, i.e., the entropy is constant, we recover the expression Pu? = constant. It was stated that this expression applied to a reversible, adiabatie process. We now see, through use of the second law, a deeper meaning to the expression, and to the concept of a reversible adiabatic process, in that both are characteristics of a constant entropy, or isentropic, process. Muddy Points Why do you rewrite the entropy change in terms of Pu? (MP 5.4) What is the difference between isentropic and adiabatic? (MP 5.5) 5.5 Calculation of Entropy Change in Some Basic Processes 1. Heat transfer from, or to, a heat reservoir, A heat reservoir (Figure 5.3) is a constant temperature heat source or sink, Because the ‘temperature is uniform, there is no heat transfer across a finite temperature differet and the heat exchange is reversible. From the definition of entropy (4S = dQrev/T), Q AS= 7 where Q is the heat into the reservoir (defined here as positive if heat flows into the reservoir.) 2. Heat transfer between two heat reservoirs ‘The entropy change of the two reservoirs in Figure 5.4 is the sum of the entropy change of each. If the high temperature reservoir is at Tyy and the low temperature reservoir is at Ty, the total entropy change is 2)+(2)-728 as=(=@)4(2)=--9 u-n, (a8) +) - re ‘The second law says that the entropy change must be equal to or greater than zero. ‘This corresponds to the statement that heat must flow from the higher temperature source to the lower temperature source. This is one of the statements of the second Jaw given in Section 5.1 Muddy Points In the single reservoir example, why can the entropy decrease? (MP 5.6) 0 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS @ LH ay Device (block of copper) Ab nochange in state a ignre 5.4: Heat transfer between two reservoirs tow Pigure 5.5: Work from a single heat re Why does the entropy of a heat reservoir change if the temperature stays the same? (MP 5.7) How can the heat transfer from or to a heat reservoir be reversible? (MP 5.8) How can AS be less than zero in any process? Doesn't entropy always increase? (MP 5.9) I Q/T =A for a reservoir, could you add Q to any size reservoir and still get the same AS? (MP 5.10) 3. Possibility of obtaining work from a single heat reservoir ‘We can rogard the process proposed in Figure 5.5 as the absorption of heat, Q, by a device or system, operating in a cycle, rejecting no heat, and producing work. The total entropy change is the sum of the change in the reservoir, the system or device, and the surroundings. ‘The entropy change of the reservoir is AS = —Q/Ty. The entropy change of the device is zero, because we are considering a complete cycle (return to initial state) and entropy is a function of state. ‘The surroundings receive work only so the entropy change of the surroundings is zero. The total entropy change is ASrorat = ASreservoir + ASaevice + ASsurroundings =-Q/Ty +040 ‘The total entropy change in the proposed process is thus less than zero, AStorat <0, CALCULATION OF ENTROPY CHANGE IN SOME BASIC PROCESSES 111 Ty Tw Cr Tr Ty-at Tear Tw. (6) Temperature equalization of (b) Reserwirs used in revenible state transforma {to biekes tions Figure 5.6: The “Hot Brick” Problem which is not possible. The second law thus tells us that we eannot get work from a single reservoir only. ‘The “only” is important; it means without any other changes occurring. This is the other statement of the second law we saw in Section 5.1 Muddy Points What is the difference between the isothermal expansion of a piston and the (Forbidden) production of work using a single reservoir? (MP 5.11) For the “work from a single heat reservoir” example, how do we know there is 90 ASsure? (MP 5.12) How does a cycle produce zero AS? | thought that the whole thing about cycles was an entropy that the designers try to minimize. (MP 5.13) 4. Entropy ehanges in the “hot brick problem” ‘We can examine in a more quantitative manner the changes that occurred when we put the two bricks together, as depicted in Figure 5.6(a). The process by which the two bricks come to the same temperature is not a reversible one, so we need to devise a reversible path. ‘To do this imagine a large number of heat reservoirs at varying temperatures spanning the range Ty — dT,...,T, +d, as in Figure 5.6(b). The bricks are put in contact with then sequentially to raise the temperature of one and lower the temperature of the other in a reversible manner. ‘The heat exchange at any of these steps is dQ = Caf’. For the high temperature brick, the entropy change is: T™ Cat Tu Shoe vee -[ oa? = Cin (®). rank = fp ir where C is the heat capacity of the brick (J/kg). ‘This quantity is less than zero. For the cold brick, AScous wick = = =Ch Ts The entropy change of the two bricks is ASivias = C fs (7) +n ( TH ‘The process is not reversible. )>e 12 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS Difference between the free expansion and the reversible isothermal expansion of an ideal as ‘The essential difference between the free expansion in an insulated enclosure and the reversible isothermal expansion of an ideal gas can also be captured clearly in terms of entropy changes. For a state change from initial volume and temperature Vi, Ty to final volume and (the same) temperature V2, T1 the entropy change is 2 2qU [2 Pav fo-f F+f . or, making use of the equation of state and the fact that dU = 0 for an isothermal process, as ‘This is the entropy change that occurs for the free expansion as well as for the isother- mal reversible expansion processes ~ entropy changes are state changes and the two system final and end states are the same for both processes. For the free expansion: ASzystem #) 3 ASsurroundings There is no change in the entropy of the surroundings because there is no interaction between the system and the surroundings. The total entropy change is therefore, 1% ASworat = ASiystem + ASaurroundings = Nn (#) >o. There are several points to note from this result (a) AStora1 > 0 80 the process is not reversible. (b) ASwatem > f2dQ/T = 0: the equality between AS and dQ/T is only for a reversible process. (c) There is a direct connection between the work needed to restore the system to the original state and the entropy change: ‘The quanti wsical meaning as “lost work” in the sense of work which we lost the opportunity to utilize. We will make this connection stronger in Chapter 6: For the reversible isothermal expansion: stom is the same as ‘The entropy is a state variable so the entropy change of the before. In this case, however, heat is transferred to the system from the surroundings (Qsurroundings <0) 80 that ASvurroundings = Qsurroundings/T <0. 5.6, MUDDIES ‘T POINTS ON CHAPTER ?? 3 The heat transferred from the surroundings, however, is equal to the heat received by the system: Quurroundings = Quystem = W. ASproentngs = Pearson n plus surroundings) is therefore Qa ‘The total change in entropy (syst ASio = ASsystem + ASwurroundings ‘The reversible process has zero total change in entropy. Muddy Points On the example of free expansion versus isothermal expansion, how do we know that the pressure and volume ratios are the same? We know for each that Py > Pi and Va > Vi (MP 5.14) Where did ASiystem = NRIn (#) come from? (MP 5.15) 5.6 Muddiest Points on Chapter 5 MP 5.1 Why is dU = TdS — PaV always true? This is a relation between state variables. As such itis not path dependent, only depends on the initial and final states, and thus must hold no matter how we transition from initial state to final state. What isnot always true, and what holds only for reversible processes, are the relations Tds = dq and Pdv = dw. One example of this is the free expansion where dq = dw = 0, but where the quantities Ts and Pdtv (and the integrals of these quantities) are not zero. MP 5.2 What makes (Quon different than AQ? The term dQrey denotes the heat exchange during a reversible process. We use the notation dQ 10 denote heat exchange during any process, not necessarily reversible. The distinction between the two is important for the reason given above in Section 5.3 MP5.3 What happens when all the energy in the universe is uniformly spread, i.e., entropy at a ‘maximum? quote from The Refrigerator and the Universe, by Goldstein and Goldstein. The entropy of the universe is not yet at its maximum possible value and it seems to be increasing all the time. Looking forward to the future, Kelvin and Clausius foresaw a time when the maximum possible entropy would be reached and the universe would be at equilibrium forever afterward; at this point, a state called the “heat death” of the universe, nothing would happen forever after The book also gives comments on the inevitability of this fate. 4 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS MP5.4 Why do you rewrite the entropy change in terms of Pv? We have discussed the representation of thermodynamic changes in P-v coordinates a number of times and it is familiar, as is the idea of the “Pv = constant” process. Iwant to relate this to the more general expression involving the entropy change Equation (5.5) to show (i) when the simple form applied and (ii) how valid an approximation it was. Using the entropy change, we now have a ‘quantitative metric for doing just that. MP 5.5 What is the difference between isentropic and adiabatic? Isentropic means no change in entropy (dS = 0). An adiabatic process is a process with no heat transfer (dQ = 0). We defined for reversible processes TdS = dQ. So generally an adiabatic process is not necessarily isentropic — only if the process is reversible and adiabatic we can call it isentropic. For example a real compressor can be assumed adiabatic but is operating with losses. Due to the losses the compression is irreversible, Thus the compression is not isentropic, MP ‘5.6 In the single reservoir example, why can the entropy decrease? When we looked at the single reservoir, our “system” was the reservoir itself. The example I did in class had heat leaving the reservoir, so that Q was negative. Thus the entropy change of the reservoir is also negative, The second law, however, guarantees that there is a positive change in entropy somewhere else in the surroundings that will be as large, or larger. than this decrease. MP 5.7 Why does the entropy of a heat reservoir change if the temperature stays the same? A heat reservoir is an idealization (like an ideal gas, a rigid body, an inviscid fluid, a discrete element mass-spring-damper system). The basic idea is that the heat capacity of the heat reservoir is large enough so that the transfer of heat in whatever problem we address does not appreciably alter the temperature of the reservoir. In grappling with approximations such as this itis useful to think about extreme cases. Therefore, suppose the thermal reservoir is the atmosphere. The mass of the atmosphere is roughly 10" kg (give or take an order of magnitude), Let us calculate the tem- perature rise due to the heat dumped into the atmosphere by a jet engine during a transcontinental ‘light. A large gas turbine engine might produce on the order of 100 MW of heat, so that the rise in atmospheric temperature, 8T atm for the heat transfer Q associated with a 6 hour flight is given by Mainty Tam = 6 3600 x 10° Substituting for the atmospheric mass and the specific heat gives a value for temperature change of roughly 10-™ K. To a very good approximation, we ean say that the temperature of this heat reservoir is constant and we can evaluate the entropy change of the reservoir as Q/T. MP 5.8 How can the heat transfer from or to a heat reservoir be reversible? We made the assumption that the heat reservoir is very large, and therefore itis temperature heat source or sink. Since the temperature is uniform there is no heat transfer across 4a finite temperature difference and this heat exchange is reversible. We discussed this in the second example, “Heat transfer between two heat reservoirs,” in Section 5.5 @ constant MP 5.9 How can AS be less than zero in any process? Doesn't entropy always increase? The second law says that the total entropy (system plus surroundings) always increases. (See Section 5.1). This means that either the system or the surroundings can have its entropy decrease if there is heat transfer between the two, although the sum of ail entropy changes must be positive, For an isolated system, with no heat transfer to the surroundings, the entropy must always increase. 5.6, MUDDIES ‘T POINTS ON CHAPTER ?? 5 MP 5.10 If Q/T = AS for a reservoir, could you add Q to any size reservoir and still get the same AS? Yes, as long as the system you were adding heat to fulfilled the conditions for being a reservoir. MP 5.11 What is the difference between the isothermal expansion of a piston and the (forbidden) production of work using a single reservoir? The difference is contained in the word sole in the Kelvin-Planck statement of the second law given in Section 5.1 of the notes. For the isothermal expansion the changes are: 1. The reservoir loses heat Q. 2. The system does work W’ (equal in magnitude 10 Q). 3. The system changes its volume and pressure 4. The system changes its entropy (the entropy increases by Q/T). For the “forbidden” process, 1. The reservoir loses heat Q. 2. The system does work W (= Q) and that’s all the changes that there are. I leave it 10 you to calculate the total entropy changes (system plus surroundings) that occur in the two processes. MP 5.12 For the “work from a single heat reservoir” example, how do we know there is no AScure? Our system was the heat reservoir itself. In the example we had heat leaving the reservoir, thus Q was negative and the entropy change of the reservoir was also negative. Using the second law, it is guaranteed that somewhere else in the surroundings a positive entropy change will occur that is as large or larger than the decrease of the entropy of the reservoir. MP 5.13 How does a eycle produce zero AS? I thought that the whole thing about eycles was an ‘entropy that the designers try to minimize. The change in entropy during a cycle is zero because we are considering a complete eycle (re- turning to inital state) and entropy is.a function of state (holds for ideal and real cycles!) The entropy you are referring to is entropy that is generated in the components of a nonideal cycle. For example in a real jet engine we have a non-ideal compressor. a non-ideal combustor and also a non-ideal turbine. All these components operate with some loss and generate entropy — this is the entropy that the designers try to minimize. Although the change in entropy during a non-ideal le is zero, the total entropy change (cycle and heat reservoirs!) is ASjorai > 0. Basically the entropy generated due to irreversibilities in the engine is additional heat rejected to the environment (10 the lower heat reservoir). We will discuss this in detail in Section 6.1 MP 5.14 On the example of free expansion versus isothermal expansion, how do we know that ‘the pressure and volume ratios are the same? We know for each that P > P, and V2 > Viv During the free expansion no work is done and no heat is transferred (insulated system). Thus the internal energy stays constant and so does the temperature. This means that P,\V, = PV; holds also for the free expansion and that the pressure and volume ratios are the same when comparing free expansion to reversible isothermal expansion. 116 CHAPTER 5. THE SECOND LAW OF THERMODYNAMICS MP 85.15 Where did ASsy,tem = NB.In(V2/Vi) come from? We were using the 1* and 2"! law combined (Gibbs) and in the example discussed there was no change in internal energy (dU = 0). If we then integrate dS = PdV/T using P/T = NR/V (with N being the number of moles of gas in volume V and R. being the universal gas constant) we obtain ASiystem = NR In(V2/V;). Chapter 6 Applications of the Second Law [vn Chapter 6; VWBE&S-8.1, 8.2, 8.5, 8.6, 8.7, 8.8, 9.6] 6.1 Limitations on the Work that Can be Supplied by a Heat Engine "The second law enables us to make powerful and general statements concerning the maximmm work that can be derived from any heat engine which operates in a eyele. ‘To illustrate these ideas, we use a Carnot eycle which is shown schematically in Figure 6.1 ‘The engine operates between two heat reservoirs, exchanging heat Qy with the high temperature reservoir at Ty, and Qz, with the reservoir at T;,. ‘The entropy changes of the two reservoirs are Asr=2t, Qu <0; Figure 6.1: A Carnot cycle heat engine 7 us CHAPTER 6, APPLICATIONS OF THE SECOND LAW to the engines by subscript “e”, Que =-Qui Qe = -Q. ‘The total entropy change during any operation of the engine is ASwa= ASy + ASy Reservoirat Ty Reservoir at Ti, For a cyclic process, the third of these (AS,)is zero, and thus (remembering that Qu <0), Qu, & ASio me Ta | Ty, Sy + AS; (6.1) For the engine we can write the first law as AU, = 0 (cyclic process) = Qre + Qe — We. We = Que+ Que Qu - Qu. ‘The work of the engine can be expressed in terms of the heat received by the engine as w= 0ne[1=(2)] a5 " ‘The upper limit of work that can be done occurs duri total entropy change ( ASiqeal) is zero. In this situation: ig a reversible eyele, for which the Maximum work for an engine working between Tyy and Ty: We = Que pi - (7 JI ; Also, for a reversible cycle of the engine, Qu , O& e+ Tw Ts ‘These constraints apply to all reversible heat engines operating between fixed temperatures. ‘The thermal efficiency of the engine is _ Work done _ W "> Feat received ~ Que 6.2, THE THERMODYNAMIC TEMPERATURE SCALE 9 ‘The Carnot efficiency is thus the maximum efficiency that can occur in an engine working between two given temperatures We can approach this last point in another way. The engine work is given by , Tr TyDSucai+ Quiz W We=-Qu or, TASworat ‘The total entropy change can be written in terms of the Carnot eycle elicieney and the ratio of the work done to the heat absorbed by the engine. The latter is the eflicieney of any eyele we can devise: We | Que. ‘The second law says that the total entropy change is equal to or greater than zero. This means that the Carnot eycle efficiency is equal to or greater than the efficieney for any other cycle, with the total equality only occurring if AStotai = 0. ASworat Muddy Points So, do we lose the capabi increases? (MP 6.1) ‘Why do we study cycles starting with the Carnot cycle? Is it because it is easier to work with? (MP 6.2) to do work when we have an irreversible process and entropy 6.2 The Thermodynamic Temperature Scale ‘The considerations of Carnot cycles in this section have not mentioned the working medium. ‘They are thus not limited to an ideal gas and hold for Carnot cycles with any medium. Earlier we derived the Carnot efficiency with an ideal gas as a medium and the temperature definition used in the ideal gas equation was not essential to the thermodynamic arguments. ‘More specifically, we can define a thermodynamic temperature scale that is independent of the working medium. ‘To see this, consider the situation shown below in Figure 6.2 which has three reversible cycles. ‘There is a high temperature heat reservoir at T, and a low temperature heat reservoir at Ty. For any two temperatures Tj, T2, the ratio of the magnitudes of the heat absorbed and rejected in a Carnot cycle has the same value for all systems. We choose the cycles 80 Qt For a Carnot cycle the same for A and C. Also Qs is the same for B and C, =1+4 64 = F(L1, Tn); 1 is only a function of temperature. Alsc ° a Qo F(T, Te), 120 CHAPTER 6. APPLICATIONS OF THE SECOND LAW Figure 6.2: Arrangement of heat engines to demonstrate the thermodynamic temperature seale =F, F(T, Ts). But 1 _ 01% Qs Q2Qs Heace FL, Ts) F(t, T) x F(T, Ts) - a OO Not function of Ts Cannot be a function of We thus conclude that F(Z, 72) has the form f(Z})/f(72), and similarly F(L2, Ts) ‘F(Z2)/f (13). The ratio of the heat exchanged is therefore Q_ pep py fD Os FT, 1) = FR)" In general, Qu _ i(Tn) Qu FT)" so that the ratio of the heat exchanged is a fimetion of the temperature. We could choose any function that is monotonie, and one choice is the simplest: f(7) = T. This is the thermodynamic scale of temperature, Qx/Qr = T/T;. The temperature defined in this ‘manner is the same as that for the ideal gas; the thermodynamic temperature scale and the ‘deal gas scale are equivalent. 6.3. REPRESEN TION OF THERMODYNAMIC PROCESSES IN T-§ COORDINATES121 Isothermal th » 3 (Mtn n+ 4 © Figure 6.3: Carnot cycle in T-s coordinates 6.3 Representation of Thermodynamic Processes in T-s co- ordinates It is often useful to plot the thermodynamic state transitions and the cycles in terms of temperature (or enthalpy) and entropy, T, S, rather than P, V. The maximum temperature is often the constraint on the process and the enthalpy changes show the work done or heat received directly, so that plotting in terms of these variables provides insight into the process. A Carnot cycle is shown below in these coordinates, in which it is a rectangle, with two horizontal, constant temperature legs. ‘The other two legs are reversible and adiabatic, hence isentropic (48 = dQeew/T = 0), and therefore vertical in T-s coordinate: If the cycle is traversed clockwise, the heat added WAS. 6 Heat added: Qy -f[ TdS = Ty(Ss— Se ‘The heat rejected (from c to d) has magnitude |Q;| = T, AS. ‘The work done by the cycle can be found using the first law for a reversible process: du =dQ-aw rds — dW (Thi form is only true for a reversible process). We can integrate this last expression around the closed path traced out by the eycle: fw = fires - f aw. However dU is an exact differential and its integral around a closed contour is zero: pris - faw. The work done by the eycle, which is represented by the term f dW, is equal to TUS, the area enclosed by the closed contour in the T-S plane, This area represents the difference between the heat absorbed ([ TdS at the high temperature) and the heat rejected (f Tas at the low temperature). Finding the work done through evaluation of § Tas is an alternative 0 122 CHAPTER 6, APPLICATIONS OF THE SECOND LAW to computation of the work in a reversible eycle from f PdV. Finally, although we have carried out the discussion in terms of the entropy, S, all of the arguments carry over to the specific entropy, s; the work of the reversible eycle per en by $ Tas. Muddy Points How does one interpret /-s diagrams? (MP 6.3) Is it always OK to “switch” ’-s and h-s diagram? (MP 6.4) ‘What is the best way to become comfortable with T-s diagrams? (MP 6.5) What is a reversible adiabat physically? (MP 6.6) 6.4 Brayton Cycle in T-s Coordinates ‘The Brayton cycle has two reversible adiabatic (ie., isentropic) legs and two reve! constant pressure heat exchange legs. The former are vertical, but we need to define the shape of the latter. For an ideal gas, changes in specific enthalpy are related to changes in temperature by dh = cd’, so the shape of the cycle in an h-s plane is the same as in a T-s plane, with a scale factor of ep between the two. ‘This suggests that a place to start is with ‘the combined first and second law, which relates changes in enthalpy, entropy, and pressure: a= Tas, ? On constant pressure eurves dP = 0 and dh = Tds. The quantity desired is the derivative of temperature, T, with respect to entropy, s, at constant pressure: (07/08)y. From the combined first: and second law, and the relation between dit and d7, this is (x), = ze (62) ant pressure legs of the Brayton eycle on a T-s plane. For a given ideal gas (specific ¢,) the slope is positive and increases as T. We can also plot the Brayton eycle in an h-s plane. ‘This has advantages because changes in enthalpy direetly show the work of the compressor and turbine and the heat added and rejected. The slope of the constant pressure legs in the h-s plane is (0h/08)y =. Note that the similarity in the shapes of the eycles in T-s and h-s planes is true for ideal gases only. As we will see when we examine two-phase cycles, the shapes look quite different in these two planes when the medium is not an ideal gas. Plotting the cycle in T-s coordinates also allows another way to address the evaluation of the Brayton cycle efficiency which gives insight into the relations between Carnot cycle efficiency and efficiency of other cycles. As shown in Figure 6.4, we can break up the Brayton cycle into many small Carnot cycles. The “i” Carnot cycle has an efficiency of where the indicated lower temperature is the heat rejection temperature for that elementar, cycle and the higher temperature is the heat absorption temperature for that cycle. The ‘The derivative is the slope of the e« 6.4, BRAYTON CYCLE IN '.S COORDINATES Figure 6.4: Ideal Brayton cy Figure 6.5: Arbitrary cycle operating between upper and lower curves of the Brayton cycle, however, have constant pressure. All of the elementary Carmot eyeles therefore have the same pressure ratio: PR = constant (the same for all cycles). From the isentropic relations for an ileal gas, we know that pressure ratio, PR, and tem- perature ratio, TR, are related by: PRO-)/7 =TR. ‘The temperature ratios (Tow /Thighs) Of any elementary eyele “i” are therefore the sane and each of the elementary cycles has the same thermal efficiency. We only need to find the temperature ratio across any one of the eycles to find what the efficieney is. We know that the temperatu st elementary eycle is the ratio of compressor exit temperature for an aircraft engine) temperature, T2/Ty in Figure 6.4. If jentary eycles has this value, the efficieney of the overall Brayton cycle (which is composed of the elementary cycles) must also have this value. ‘Thus, as previously, Tintor easton = 1 — . Feompresor exit A benefit of this view of efficiency is that it allows us a way to comment on the efficiency of any thermodynamic eycle. Consider the eycle shown in Figure 6.5, which operates between 124 CHAPTER 6, APPLICATIONS OF THE SECOND LAW some maximum and minimum temperatures. We can break it up into small Carnot cycles and evaluate the efficiency of each. It can be seen that the efficiency of any of the small cycles drawn will be less than the efficiency of a Carnot cycle between Tiage and Train. T graphical argun efficiency of any other thermodynamic cycle operating between these maximum and minimum temperatures has an efficiency less than that of a Camot eyele. Muddy Points If there is an ideal efficiency for all cycles, is there a maximum work or maximum power for all cycles? (MP 6.7) 6.4.1 Net work per unit mass flow in a Brayton cycle In Section 3.7.1 we found the net work of a Brayton cycle in terms of heat transfer. Now that we have defined entropy, we can reexamine the net work using an enthalpy-entropy (Is) diagram, Figure 6.6, The net mechanical work of the cycle is given by: Net mechanical work/unit mass = Purbine ~ Weompressors where, by the first law, Mhcomp Wurbine = Ahas = —Ahiarb: Weomprensor = — A If kinetic energy changes across the compressor and turbine are neglected, the temperature ratio, TR, across the compressor and turbine is related to the enthalpy changes: Alicomp _ [Ato] eae = ‘The net work is thus ‘The turbine work is greater than the work needed to drive the compressor, as is evident on the (h-s) diagram. 6.5 Irreversibility, Entropy Changes, and “Lost Work” Consider a system in contact with a heat reservoir during a reversible process. If there is heat @ absorbed by the reservoir at temperature 7, the change in entropy of the reservoir is AS = Q/T. In general, reversible processes are accompanied by heat exchanges that occur at different temperatures. To analyze these, we can visualize a sequence of heat reservoirs at different temperatures so that during any infinitesimal portion of the cycle there will not be any heat transferred over a finite temperature difference. 6.5. IRREVERSIBILITY, ENTROPY CHANGES, AND “LOST WORK” Figure 6.6: Brayton cycle in enthalpy-entropy (h-s) representation showing compressor and turbine work a ‘R Figure 6.7: Irreversible and reversible state changes During any infinitesimal portion, heat dey will be transferred between the one of the reservoirs which is at T. If dQry is absorbed by the system, the entropy change of the system is Que Seystem = + ‘The entropy change of the reservoir is Qe AS eerie = Fe ‘The total entropy change of system plus surroundings is AS erat = ASzyetem + ASreservoir = 0- ‘This is also true if there is a quantity of heat rejected by the syst ‘The conclusion is that for a reversible process, no change occurs in the total entropy produced, i., the entropy of the system plus the entropy of the surroundings: AS\ouat = 0. We now carry out the same type of analysis for an irreversible process, which takes the system between the same specified states as in the reversible process. ‘This is shown schematically in Figure 6.7, with T and R denoting the irreversible and reversible processes. In the irreversible process, the system receives heat dQ and does work di. ‘The change in 126 CHAPTER 6, APPLICATIONS OF THE SECOND LAW internal energy for the irreversible process is dU =dQ— dW (Always true — first law). For the reversible process dU = TdS ~ dWrey. Because the state change is the same in the two processes (we specified that it was), the change in internal energy is the same. Equating the changes in internal energy in the above two expressions yields AQocrunt — AWaetuat = TAS — dWerew ‘The subscript “actual” refers to the actual process (which is irreversible). ‘The entropy change associated with the state change is — AQuetant | “s iS = esl [2% = dWeetuat (6.3) If the process is not reversible, we obtain less work (see IAW notes) than in a reversible process, dWcyust < dWeey, 80 that for the irreversible process, AQucrvat as > There is no equality between the entropy change dS and the quantity dQ/T for an irreversible process. The equality is only applicable for a reversible process. ‘The change in entropy for any process that leads to a transformation between an initial state “a” and a final state °b” is therefore AS = S,-S,> where dQnecue! is the heat exchanged in the actual process. The equality only applies to a reversible process, The difference dWroy ~ dWactual Tepresents work we could have obtained, but did not. It is referred to as lost work and denoted by Wigge- In terms of this quantity we can write, = dQnerst r Wiest as 7 (64) The content of Equation (6.4) is that the entropy of a system can be altered in two ways: (i) through heat exchange and (ji) through irreversibilities. ‘The lost work (dWjog, in Equation (6.4)) is always greater than zero, so the only way to decrease the entropy of a system is through heat transfer. ‘To apply the second law we consider the total entropy change (system plus surroundings). If the surroundings are a reservoir at temperature T, with which the system exchanges heat, Querust AS rservie(= AS yaroundngs) t 6.5. IRREVERSIBILITY, ENTROPY CHAD (GES, AND “LOST WORK” 127 ‘The total entropy change is AS porat = USaystem + USeursoundings = (Sex + Wow AScoea1 = > 0. ‘The quantity (dWicg./T) is the entropy generated due to irreversibility. Yet another way to state the distinction we are making is ASzystem = AStrom heat + AS generated due to = FSieat transfor + 1Scen- (6.5) tamer Freveraible proces ‘The lost work is also called dissipation and noted d. Using this notation, the infinitesimal entropy change of the system becomes: db Septem = ASpeos emnster + or TdSeysem = dQ + de Equation (6.5) can also be written as a rate equation, as a Shot tannntor + Scien (6.6) Either of Equation (6.5) or (6.6) can be interpreted to mean that the entropy of the system, S, is affected by two factors: the flow of heat Q and the appearance of additional entropy, denoted by dScen, due to irreversibility', This additional entropy is zero when the process is reversible and always positive when the process is irreversible. Thus, one can say that the system develops sources which create entropy during an irreversible process. The second law asserts that sinks of entropy are impossible in nature, which is a more graphic way of saying that dSGen and Sgen are positive definite (always greater than zero), or zero in the special case of reversible processes. The term Sheat transfor (- which is associated with heat transfer to the system, can be interpreted as a flux of entropy. ‘The boundary is crossed by heat and the ratio of this heat flux to temperature can be defined as a flux of entropy. There are no restrictions on the sign of this quantity, and we can say that this flux either contributes towards, or drains away, the system's entropy. During a reversible process, only this flux can affect the entropy of the system. This terminology suggests that we interpret entropy as a kind of weightless fluid, whose quantity is conserved (like that of matter) during a reversible process. During an irreversible process, however, “This and the following paragraph are excerpted with minor modifications from A Course én Thermody- namics, Volume 1, by J. Kestin, Hemisphere Press (1979)

You might also like