1
PHI LIP G. H ILL Deparlmem of Mechanical Engineering Massachusetts [nstitute a[Technology
CARL R. PETERSON Research and Development , Ingersoll-Rand Company Bedminster, New Jersey
MECHANICS AND THERMODYNAMICS OF PROPULSION
Menlo Park, California
ADr;>ISON.WESLEY PUBLISHING COMPANY Reading, Massachusetts Londoll' Amsterdam . 0011 Mills, Ontario' Sydney
This boo k is in the
ADDISON-WESLEY SERIES IN AEROSPACE SCIENCE HOWARD
W.
EMMONS,
S. S.
PENNER,
Consultillg Editors
Preface Th is boo k is based on an introductory co urse in propulsion which has been given for several years to students in aerona utical and mechanica l engineering at Massachuse tts Institute of Technology. It deals mainly with the propulsion of Th ird print ing, November 1970 Copyright ® 1965 Philippines Copyrighf 1965 ADDISON-WESLEY PUBLISHING COMPANY, INC.
Printed ill the United 5{(aes oj America ALL RIGHTS RESERVED. TH IS BOOK, OR PARTS THERWF' , MAY NOT ijE REPRODUCED IN ANY FOR""I
WITHOUT WRITTEN PERMISS ION OF THE PUBLISHERS
Library oj Congress Catalog Card No. 65-10408 ISBN 0-201-02838-7 RST-MA-898765 43
aircraft and space vehicles. The objective of the book is to show how basic pri nciples may be used to predict the behavior of propulsion devices. There are, of course, many methods of propulsion, but the underlying physical principles are relatively few. Often there is a large gap between theoretical studies and their useful app lication to actual engineering pr061ems. This book does not pretend to close such a gap, but it does stress th e app lication of th e student's theoreti cal knowledge to severa l real devices, poi nt ing toward the engineerin g objectives of prop ulsion. Ex perience has show n that the app lication of fundamenta ls to these problems can reward the student not only with a valuable appreciation of their practica l sign ificance but also wi t h a deeper understa nding of the principles Ihemselves. To this end , the book attem pts to show how basic physical laws bot h describe and limi t the perfor mance of particular devices. With the present trend toward increasingly basic engineering curricula there rema ins a definite need for at least one co urse in which a va riety of fundame ntal ideas are brought together and made to interact so that their engineering implications can be clearly seen. Pro pulsion is a n ad mirable subject for this purpose, since it significan tly involves so many di sci plines. This text may therefore serve a more general purpose than simply conveying knowledge of propulsion devices to the student. In general, the most sa tisfacto ry performance of a given propulsio n engine is determined not only by the laws of mechanics and thermody na mics, bUl a lso, to a large extent, by the behavior of materials. The mai n emphasis in this book is on the former, though ma te ri al limitati ons a re identified and discussed. The mathematical portions of the tex t assume a know ledge of calculus. In addition , a n introductory course in vector calculus is des irab le, but not essential. An attempt has been made to minim ize mathematical complication so that physic al principles can be more readil y di scerned , and t he proble m sect ions at the ends of iii
iv
PREFACE
the chapters are designed to give the reader an idea of the practical applications of these principles. Throughout the text much attention has been given to the "working fluid " of the propulsion engine. Generally it is considered to be a continuum, since the principal interest lies in its macroscopic behavior. Two exceptions to this rule occur. First, in the study ofpJasma, it is necessary to examine microscopic processes in order to understand and predict the macroscopic behavior. Second, in the study of ion or electron engines, it is often necessary to discuss the behavior of discrete particles. The control volume method of analysis, whose usefulness for fluid flow has been demonstrated so thoroughly by Shapiro, has been used extensively. One-d imensional approximations are frequently employed, but certain two-dimensional phenomena like heat transfer and other boundary layer behavior are also discussed. Basically, two classes of propulsion devices are considered : (1) air-breat hing engines, and (2) rockets.
Air-breathing engines have been under very intensive development over the past 25 years. They are particularly inst ructive examples of engi neering achievement, because of both the magni tude and the success of the development effort. It is most illuminating to see the power and the limitatio ns of theoretical methods and to observe the mixture of art and science in thei r evolution. Although it has now approached some degree of completion, development continues toward better performance and new applications, so that air-breathing engines will remain a source of challenging engineering problems for some time to come. Rock et engines, by comparison, are currently under much more act ive development. M uch is now known about chem ical rockets, but nuclear and electrica l rockets are only entering the experimental stage. The latter are nonetheless discussed in this text, since current developments indicate that nuclear and electrical phenomena will be important in future propulsion engines. There is need for an introductory discussion of th,e application of these phenomena to propulsion. Part I is concerned mainly with a review of those topics in thermodynamics, combustion, and fluid mecha nics which are relevant to propulsion engines. One of the purposes of th is review is to introduce necessary concepts and laws in the notatio n used in su bsequen t chapters. In Chapter 2 the thermodynamics o f equil ibrium combust ion reactions are discussed in some detai l because of their relevance to chemical rockets especially. Since propulsion systems generally deal with accelerating fluid s, the gas dynamics of isentropic expansion are reviewed in Chapter 3, a long with the modifying effects of friction at solid boundaries, and shocks. In many engines, e.g. , the turbojet, the viscous boundary layer exercises a controll ing effect on the performance of the engine; hence the question of boundary layer behavior is discussed in some detail in Chapter 4. Convect ive heat transfer is also briefly reviewed, chiefly because of its relevance to the cooling of rocket motors. Electrostatic and electromagnetic acceleration are reviewed in Chapter 5 because they are centrally important in the rising genera ti on of propulsion engines
PREFACE
v
for space flig ht. We also give an elementary description of the behavior of charged particle beams and conducting fluid s, in order to aid understanding of releva nt phenomena, though it is not entirely clear at present which phenomena will be of greatest significance in working electrical-thrust engines. Part 2 is confined to ai r- breathing engines. Chapter 6 demonstrates that the performance of any air-breathing engine can be pred icted from the laws of thermodynamics, given the maximum permissible temperature and the appropriate efficiencies of engine com ponents, e.g., the compressor. Cha pter 7 deals with t he components of the ramjet (diffuser, combustor, and nozzle) and shows how their efficiencies depend principally on aerodynamic phenomena. The essential difference between the ramjet and the turbojet is simply the turbomachinery, which is the subj ect of Chapters 8 and 9. Fortuna tely the turbojet is so highly developed lhal several interesting examples of act ual engine configurat ions and performance can be cited by way of illustration. Part 3 is confi ned to rockets. Chapter 10 presents an elementary treatment of rocket vehicle performance in order to show the significance of specific impulse and other variables, and also to ind icate the relat ive advantages of low- and highth rust engines for various missions. The trajectories of space veh icles are briefly disc ussed in order to show what they req uire of the propulsion system. Chemica l rock ets are the su bject of Chapte rs I I, 12, 13 and 14, wh ich ta ke up a series of pertinent top ics. The combustio n process, the gas dynamics of the nozzie, and the cooling of the wall are all exami ned in some deta il , as well as auxiliary equipment such as the tur bopump. Chapter 15 is concerned with the nuclear rocket, in which a nuclear reactor is used to heat a gas such as hydrogen to high temperature before it expands in th e nozzle. The limi tat ions of this engine and th e phenomena th at govern its performance are briefly discussed. Chapter 16 focuses on electrostatic and elect romagnetic rockets, although these are in ea rly stages of development. Co nsiderable experimental work has been done on the components of the electrostat ic rocket, and it appears relatively promising for future applicati on, provided suitable electrica l energy conversion mac hinery can be developed. The electroth ermal and three kinds of possible electromagnetic rockets are briefly desc ri bed. This text may be used fo r either a one- or two-term course in propulsion. For a two-term sequence, the first term could be used fo r either air- breathing or rocket engines. A o ne-te rm course could be based o n Chapters 6, 10, and II , with selected portions of Parts 2 and 3 and appropriate review materia l. The aut hors wo uld li ke to reco rcl their indebted ness to Professor Edward S. Taylor, who has been our teacher, critic, and friend during our years or associa tion in the M. l.T. Gas Tu rb ine Laboratory. The text has been written largely as a result of his encou ragement, and it may be that his influence will be recogni zable in some of the more worthwhile portions. We are certainly grateful to him for writing the introductory chapter. Throughout the text, there are a number of references to the work of Professors J. H. Keenan and Ascher H. Shapiro. Their expositions of thermodynamics
vi
PREFACE
and fluid mechanics were so enlightening to us as studen ts that we could not help drawing heavily from them in presenting t he subject of propulsio n. The preparation of ea rl ier editions of classroom notes upon which the book is based was supported in part by a grant made to the Massachusetts Institute of Technology by the Ford Foundation for th e pu rpose of aiding in the improvement of engineering education. This support is gratefully acknow ledged. The authors wou ld also like to ac knowledge the di li gent and efficient work o f Mrs. Madelyn Euv ra rd and Mrs. Rose Tedmon in piecing together many fragments and revisions of the text. Cambridge, Massachusetts October 1964
P.G .H. C.R.P.
Contents PART I
Introduction and Review 01 Fundamental Sciences
Chapter 1
PROPULSION
1- 1 1-2 ]-3 Chapter 2
2- 1 2- 2
2- 3 2- 4
Cha pter 3
3-1 3-2
3-3 3-4 3-5 3-6
Theory of propulsion Development of propul sive devices Future possibilities .
I 3 5
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
Int roduction. The laws of mechan ics and thermody nam ics fo r fluid flows Mass conservation and the continuity equation. Newton's seco nd law and the momentum eq uation. The first law of thermodynamics and the energy eq uation. The second law of thermodynamics and the entropy equation Thermodynamics of gases . Equilibrium combustion thermodyn amics; chemical reactions Mixtures of gases. Chemical transformation. Composition of products
7 8
]5 20
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
Introduction. Genera l one-dimensional flo w of a perfect gas Stagnation state. Mach num ber Isentropic flow . Frictionless co nsta nt-area flow wit h stagnation temperatu re change Consta nt-a rea flow with fr iction
Shocks
43 43
47 49 53 57
Normal shocks. Oblique shocks Chapter 4
4- 1 4-2 4- 3 4-4 4-5
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
The viscous boundary laye r The bounda ry layer equations. La minar boundary layer solutions The turbu lent boundary layer . Boundary layer heat tra nsfer vii
64 71 76
82 91
viii
CONTENTS
CONTENTS
Chapter 5
5-1 5-2
5- 3 5-4 5-5
5- 6 5-7
ELECTROSTATICS, MAGNETO STATICS, AND PLASMAS
Introduction . Basic equation s of electrostatics and magnetostatics Motion within electric and magnetic fie lds Plas mas Conducti on in plasmas in the absence of magnetic fields Conduction and momentum exchange in plasmas with magnetic fields One.dimen sional steady flow of a plasma Constant·area , constanHemperature acceleration. Magnetic Reynolds number. Plasma boundaries
PART 2
Air.Breathing Engines
Chapter 6
THERMODYNAMICS OF AIRCRAFT JET ENGINES
6- 1 6-2
6-3 6-4
6-5 Chapter 7
7- 1
7- 2 7- 3 7-4
7-5 Chapter 8
8- 1 8- 2
8- 3 8-4 8- 5 8- 6
Introduction. Thrust and efficiency The thrust equation. Engine performance . Aircra ft range. Take· off thrust. Specific fuel consumpt ion The ramjet The ideal ramjet. The effect of aerodynamic losses Gas turbine engi nes . The turbojet. The turbofan engine. Turboprop and turboshaft engines Typical engine performance
Chapter
102 102 106
107 112 117
123
139 140
149 157
176
AEROTHERMODYNAMICS OF INLETS, COMBUSTORS, AND NOZZLES
Introduction. Subsonic inlets and diffusers Flow patterns . Internal flow. External flow. Diffuser performance Supersonic inlets Reverse-nozzle diffuser. The starting problem. External deceleration Combustors . Combustion limits and Harne speed. Ram jet combustors and after· burners . Turbojet combustion cha mbers . Ana lysis . Alt itude per· formance Exha ust nozzles .
190 191
201 215
232
JET ENGINE TURBOMACHINES, AXIAL COMPRESSORS
Introduction . Angular momentum Single·stage axial compressors Press ure-rise limitations. Cascade aerodynami cs . Efficiency. Offdesign pe rformance. Degree of reaction. Solidity. Radia l varia tions Mult istage axial compressors . Instabilities and unsteady flow Surge. Rotating stall Aerodynamics of starting
238
239 244
9
272 275
JET ENGINE TURBOMACHINES, CENTRIFUGAL COMPRESSORS AND AXIAL TURBINES
9- 1 Int roduction. 281 9- 2 The centrifuga l compressor 281 ImpeUer and inducer. Difruser • Performan ce. Compa rison of centrifugal and ax ia l compresso rs 9- 3 The axial turbine 294 Stage dynamics . Radial variations . Deviation . Efficiency. Tur· bine performance 9- 4 Turbine and compressor matching 307 9- 5 Stresses in turbines and compressors. 310 PART 3
Rockets
Chapter 10
PERFORMANCE OF ROCKET VEHICLES
10- 1 Introduct ion. 10- 2 Static performance Thr ust. Specific impulse 10-3 Vehicle accele rat ion Gra vity . Drag. Single·stage so unding rocket. Burning time 10-4 Chemical rockets Single-stage rockets . Multistage roc kets 10- 5 Electrical rocket vehicles 10- 6 Space missions . High thrust. Low thru st Chapter 11
Chapter 12
320 322 327
335 338
353 354
CHEMICAL ROCKETS, PROPELLANTS AND COMBUSTION
Introduction Equilibrium composition Liquid propellants . Solid propellants Combustion chambers Liquid·propellant combusti on chambers . Solid· propellant tion chambers 12- 6 Combustion instabilities
12- 1 12- 2 12- 3 12- 4 12- 5
Chapter 13
319
CHEMICAL ROCKETS, AN INTRODUCTION
11- 1 Characteristics of chemical rockets 11-2 Analysis of an ideal rocket
13- J 270
ix
365 365 368 372 375 COIn bus·
392
CHEMICAL ROCKETS, EXPANSION IN NOZZLES
Rocket exhaust nozzles 397 Nozzle shape. The effect of fri ction and heat transfer. The errect of back pressure on nozzle flow. Plug and expansion-deflection nozzles. The effect of propellant property variations all nozzle flow. The effect of solid or liquid particles in the flow
x
CONTENTS
Chapter 14
CHEMICAL ROCKETS, THRUST CHAMBERS
14-1 Introduction. 14-2 Rocket heat transfer Regenerative cooling. Heat sinks. Convective film coefficients. Radiative heat -transfer 14-3 Rocket construction Liquid-propellant rockets: regenerati ve cooling systems. Solidpropellant rockets: casings. Solid-propellant rockets: nozzle~ 14-4 LIquid-propellant pressurization . Gas pressurization. Turbopumps • Centrifugal pumps. Cavitation. Turbines 14-5 Selection of combustion pressure. Liquid-propellant engines. Solid-propellant engines
14-6 .Ignition . 14-7 Thrust vectoring Chapter 15
1
t.
439
445
Introduction and Revie""
460
01 Fundamental Sciences
461 462
NUCLEAR ROCKETS
15-1 15-2 15-3 15-4
Introduction. Nuclear reactions as energy sources Nuclear reactors The solid-core nuclear rocket . Heat transfer to the rocket propellant. Rocket performance 15-5 Advanced nuclear rocket concepts
Chapt.r 16
PART
425 425
468 469 475 480 486
ELECTRICAL ROCKET PROPULSION
16-1 Introduction. 16-2 Electrostatic propellant acceleration . 16-3 Sources of charged particles . Bombardment ionization. Contact ionization
490 491 495
16-4 The plane diode
506
16-5 Beam optics. Exhaust neutralization
511
16-6 The arejet
518
16-7 Stea9Y crossed-field accelerators 16-8 Pulsed-plasma accelerators 16- 9 Traveling-wave accelerators
522
527 528
Append;x I
OPTIMIZATION OF MULTISTAGE ROCKETS
535
Appendix II
TABLES.
539
Appendix III
LIST OF SYMBOLS
545
Answers to Selected Problems
549
Index
553
.-
Kitty Hawk, North Carolina, 1903: man's first plane flight. The Wright brothers' famous flying machine was the first plane to lift a man into the air. Orville Wright took the first airplane ride, lying face downward on Ihe lower wing or Ihe biplane. His ride, which was only a rew feet above the ground, lasted 12 seconds and covered 120 reet before running into the ground. Photograph courtesy the Library of Congress and United Airlines.
1 Propulsion*
1-1
THEORY OF PROPULSION
The verb "propel" means to dr ive or push forward or onward. Clearly th is defin ition implies the existence OWl body. a force (drive or push). a nd a prefem;d di rection (forward or onward). .tccording to Newton's third law of motion, forces occur only in equal and opposlle pa irs, and the existence 01 a force on a body therefore entails a reaction. This reaction cannot be on the body itself, since a pair of equal and opposite forces on the same body wi ll not "drive or push forward or onward." Therefore one or more additional bodies are esse.u.tial...tQ.. propulsion. For a self-propelled vehicle like an automobile, the second body is Ihe earth aself, and the act of propulsion entails a transfer of momentum between the automobile and the earth. In cases of particular interest in this text, the second body is usually a fluid med ium. In the instance of a propeller-driven a irplane, this fluid is the surrounding air, while in the case of a rocket it is a fluid carried along with, and discharged from, the rocket itself. In a jet ajmlane. the major part of the fluid used in propulsion is the surroundin atmos here but the fuel, which ..Qlay canslltu e 0...'> e PAQOl! sive fluid , is carried along and discharged as In a rocket. If ; consider an airplane in st raight level flight at uniform speed, it is clear that the horizontal component of the resultant force acting on the airplane must be zero, in order to satisfy Newton 's second law of mo tion. We know that to propel the airplane it is necessary to overcome the drag (frictional and induced) of the fluid on the body. If this is done by means of a propeller, the fo rce of the propeller on the airplane must be sufficient to cancel the drag forces . .}he propeller can produce this fore onl b transferrin momentum to the strea m orair which passes througlLit. The drag forces also transfer the momentum to the surroun mg air; and indeed, the horizontal momentum change due to the propeller thrust and the horizontal momentum change due to drag forces must be equal and opposite, I
* An introduction by Edward S. Taylor, Professor of Flight Propulsion, Department of Ae ronautics and Astronautics, Massac hu sett s Institute of Technology. 1
p
1-1
1-2
whet her we are speaking of the momentum of the airplane or of the fluid medium surrounding it. Thus, although the air is disturbed by the passage of such an
1-2
airplane, there is no net c hange in its horizontal momentum before and after the
The history of propulsion illustrates the interaction between economic or military need and technological advances. Although da Vinci's sketches of screw propellers date from the fifteenth (or possibly early sixteenth) century, the need for such a device did not arise until the development of the steam engine provided an appropriate dri ving mechanism . The first successful steamboats were driven by paddle wheels, but the paddle wheel must operate at the interface of two media (air and water) and is therefore clearly unsuitable for submarines, lighter-than-air ships, and airplanes. (It might also be said to be unsuitable for any ship when the sea is ro ugh and the in terface is shifting a bouL) The first successful submarine, the first airs hip, and the first airplane all had screw propellers. The airplane propeller still serves well when the speed of the aircraft is less than 350 mi les per hour. When man began to want to fly at higher speeds (originally for military reasons), a lighter and smaller power plant was essential. The jet engine answered thi need, while at the same time avoiding the difficulties encountered by propellers f operatin g at supersonic tip velocit ies. It is interesting to note that there was no incentive for developing a jet engine until the art of airplane design had advanced enough to make high-speed flight practical. As a matter of fact, the jet is inappropriate to low-speed flight, and the propeller provided an essential first step without which we would almost certainly not have developed airplanes at al l. An interesting history of the development of aircraft engines is to be found in Reference 4. The jet engine is a good example of the solution of a technical problem-that of high-speed flight-which required that many conditions exist simultaneously. First the need had to exist; then there had to be sufficient mastery of the art of
2
PROPULSION
passage. If we consider an accelerating rocket (assumed, for the purpose of simplifying the argument, to be in gravity-free flight outside the atmosphere), the total momentum of the rocket and the material which it has discharged must remain constant. It is inconvenient to keep track of the momentum of the propellant after it leaves the rocket, and we shall see (Chapter 2) that in order to calculate acceleration of the rocket, it is necessary only to observe the momentum of the jet leaving the rocket at the nozzle discharge. Clearly the principles of propulsion are merely applications of Newton's laws of motion. Long before Newton announced his laws of motion (1687) which now make the principles of propulsion comprehensible to us, devices for elTeeling propulsion were known and used. The Chinese had roc kets as early as the twelfth century, and screw propellers are visible in some of Leonardo da Vi nci's dra wings. About 200 B.C., the Egyptian philosopher Hero invented a nd constructed a turbine which clearly shows the principle of jet propulsion. In pro pulsion, as in most other technological adva nces, the art preceded the science. However, it is highly improbable that man could ever have produced supersonic ai rcraft or rockets capable of pUlling vehicles in orbit aro und the eart h withou t a very good theory of propulsion. By "theory" we mean a method of generalizing the results of experiments so that we can, with some confidence, predict the results of new ex periments.
A good theory enables one to see the main points without getting confused by details, and makes it possible to perform limit analysis; e.g., to determine what could be done if the machinery could be made without losses. F urthermore, a theory facilitates communication between workers. Moreover, even poor theory
may stimulate thought and experiment which will lead to new ways of accom plishing things. A good theory can prevent wasting elTort on unworkable devices. Many erroneous ideas of propulsion by jets were prevalent as late as the 1920's. It was well known at this time that a fi reboat di recting its streams of water at a fire would gradua lly move away from the fire. Eminent professors of engineerin g
ha ve been heard to scoff at this observation, say in g that "you can't push on a
rope." This statement, of course, reflects a profound ignorance of the problem, which has nothing to do with whether the streams fro m the fire nozzle ever reach any target. Goddard, at about this time, found it necessary to demonstrate that his rockets had thrust even when fired in a vac uum. T his points up the prevalence of the same fallacy as ex isted in the fireboat problem. It apparently was difficult for people to conce ive of something propelling itself without some med ium or body to push on. The id ea of the device pushing on a med ium which was being discarded was
DEVELOPMENT OF PROPULSIVE DEVICES
3
DEVELOPMENT OF PROPULSIVE DEVICES
airplane design, including aerodynamics, structure, and control; and fin ally there
had to be the knowledge of turbomachinery design and the availability of materials capable of withstanding high temperatures. These are only some of the major hurdles. Of course the solution of the problem also depended on a sophisticated manufact uring technique-forging, casting, co ld-forming, welding, and machining - and on the availability of countless "minor" but essential items: bearings, seals, lUbricants, pum ps, control devices, etc. In any development as complex as the jet engine, special circumstances are bound to arise which require changes and advancement of the technology in these "minor" items. However, without a rea-
sonable state of development in these as well as the major items, the development period would be likely to stretch out until it became clear that the project had better be postponed until the necessary supporting technology was able to catch up. John Barber patented a gas turbine in 179 1. There is little doubt that Barber had insufficient knowledge of basic mechanics, and certa inly insufficient knowledge
understood by only a few. Now it is a part of the knowledge of most if not all
of fluid mechanics to develop a successful gas turbine; but even if he had had all the fundamental knowledge we have today, his job would still have been impossible unt il the general technological level had risen enormously. Materials and manu-
newly graduated mechanical and aeronautical engineers.
factu ring techniques were not adequate even to begin this task until well over a
p
4
1-2
PROPULSION
hundred years later. Furthermore, had he been able to build a gas turbine, there would have been no economically feasib le use for it. . History is full of such examples of inventions made too soon. They emphaSIZe the importance of the engineering task of determin ing when a need eXIsts and when
supporting technology, both in theory and practice, are sufficiently advanced to begin a new deve lopment. To determine these facts involves a knowledge or at least an estimate of what extensions of theory and what refinements of practIce are essential to its success.
The first successful aircraft propulsion device was Henry Gifford's steamengine, propeller-driven di ri gible balloon, which went from Paris to Trappes in 1851 at an average speed of about seven miles per hour. However, for an aIrplane power plant, the necessary water supply (or condenser) and boiler would be so unreasonably heavy that even if the steam engine weighed n~thi n g, o.n ly very P?or airplane performance would be possible. The successful aIr plane, III a practIcal sense, had to wait for the development of the internal combustIOn engIne, whIch eli minated the necessity for either a wate r supply (or condenser) or a bolier. At
FUTURE POSSIBILITIES .
1-3
5
While the German V-2 rocket, with a 2000-pound warhead and a range of about a hundred miles, had considerable value as a morale-bu il der for the Germans, the damage done by these rockets was minor and the accuracy was poor compared with the effort invo lved. However, with the advent of nuclear devices, the potent ia l explosive power of a warhead of given size was mu ltiplied by an eno rmous factor, and the effectiveness of a long-range missile with even poor accuracy became
enormous. T his, then , was the motivating force behind the development of rocket propulsion ; it was inevi ta ble that billions of dollars would be poured into the development of such dev ices. Except for the guidance devices (where enormous advantages were yet to be obtained by subsequent improvements), the art of rocket ry was reasonably ready and the science had been ready for some time, with the exception of one factor: the reentry of a rocket a t high speed. This problem was not sufficiently understood until much later, but practically speaking it proved to be less difficult than was feared. Thus in the forties the need was evident and the tec hnology and science existed. With reasonab le luck and not-unreasonable improvement in the technology, the job could be done.
the time of the Wright brothers' experi ments, available internal combustion
engines were still too heavy, but they did exist, a nd the time was ri pe fo r their development to a reaso nable engine weight per horsepower. The Wnght brothers had to take on this problem [1 J, along with the design and constructIOn of the first powered airplanes to demonstrate controlled flight. . Langley, who was faCIng the same problem at the same time, found an extraord inary englllcer
In
Manley [2],
1-3
FUTURE POSSIBILITIES
At the present writing (1964) we see successful chemical rockets, and the need, if we are to travel far in the solar system, for vastly improved means of propulsion.
The possibili ties now appear to be: (I) Improvement in chemical rockets, which undoubtedly will take place, but
who handled this pa rt of the problem with outsta nding success. The possibility of propulsion by means of a jet had been well known [3J for many years before it was usefully applied by von Ohain (first flight August 27, 1939) and Whittle (first fli ght May 15, 1941). Before jet propulsion could become feasible, it was essential tha t ai rplanes be des igned and built for high speed, and also it was necessary that the art of design and construction of high- temperature
(2) Nuclear rockets . These hinge on our ability to make nuclear reactors which will operate at very muc h higher temperatures than they do at present. (3) Some form of electrical propU lsion. Th is appears to req uire better understanding of the properties and behavior of high-temperature ionized gases, and
turbines and efficient compressors be advanced cons iderably to meet the stringent
considera ble improvement in the means for producing electric power in space.
may be insufficient to an swer the needs for ex ploring the solar system.
requirements (primarily light weight combined with high en1ciency) of a useful
What route we shall take from here is unclear. It may be that improvement in
aircraft engine. The development of the sophist icated rocket is an illustration of a similar series
conven tiona l chemica l rockets wi ll offer the least expensive alternati ve. If we ca n judge from history, it may even be that we shall wait for new scientific discoveries J new materials, and new improvement s in the technology of ma nu factur ing before the task can be accomplished. In any case we must rem ember that a continuing
of events. The Chinese seem to have used the rocket mostly for fireworks, although there is some indication of military use.
In the early days of artillery, acc uracy
was so poor that even a crude unguided rocket could compete, and about the beginning of the nineteenth century the British officer Congreve was instrumental in develop in g rockets as a subs titute for ar tillery. There is no doubt that these rockets were useful in frightening cavalry horses, and perhaps were not much less accurate than the artillery of the time. The "rocket's red glare" of the national anthem refers to Congreve rockets in the War of 1812. Subseq uent ly artillery was greatly improved and military rockets became obsolete. Rockets, in order to be effective, would have to be guided. The necessary understand ino of con trol problem s came about in the 1930's, and th e requisite development ~f th is technology came soon after.
strong desire to do the job will be essential to its accomplishment.
F 6
PROPULSION
References 1. MEYER, R. B., JR., Annual Report of the Smithsonian Institutio n, 1961 2. "A Description of the Manley Engine," Soc . oj Automotive Engineers, 1942 3. SCHLATFER, R. , and S. D. HERON, De velopment oj Aircraft Engines m,ld FI~els. Cam-
2
bn'dge , Mass ( .: Graduate School of Business Administration, Harvard Umverslty, . . I1950.
4. T AYLOR, C. F., "Aircraft Propulsion." Ann ual Report of the Sm ithsonIan nshtu-
Mechanics and
tion, 1962, pages 245-298
Thermodynamics of Fluid Flo""
2-1
INTRODUCTION
An understanding of fluid mechanics and thermodynamics is perhaps the most important prerequisite for the study of propulsion. In nea rly all the propulsion methods discussed in this volume, thrust is developed by imparting momentum to fluid streams. Most of the methods involve thermal effects in one way or another. T he purpose of this chapter is to set forth a concise statement of the principles of mass, momentum, energy, and entropy in a for m which is suitable to the treatment of fluid streams. We shall ass ume that the student is th oroughly fa miliar with the first and second laws of thermodynamics and their corollaries as applied to ord inary nonreacting continuum substances in equilibrium. Hence, after careful definition of the concepts of system and control volume, these laws will merely be stated in fo rms which the reader will find useful in the remai nder of the text. The thermodynamics of equilibrium chemical reactions will then be treated in some detail, because of its relevance to combustors in propulsion engines. Considerable literature is ava ilab le as background material for the subject matter of both this chapter and Chapter 3. This summary largely follows the material of Keenan [I] and Shapiro [2]. T he concepts of syslem and cOlllrol volume, as applied in the following pages and throughout the text, have very specific meanings. The purpose of these concepts is the specification either of a defin ite collection of material or of a region in space which is to be analyzed. System . A system is a collect ion of matter of fixed identity. It may be considered enclosed by an invisible, massless, flex ible surface through which no matter can pass. The boundary of the system may change position , size, and shape. 7
8
MECHANICS AND THERMODYNAMICS OF FLU ID FLOW
2-2
COlltrol vO/llme. A control volume is a region of co nstant shape and size which is fixed in space relative to the observer. One can imagine an invisible, massless, rigid envelope (the control surface) enclosing the control volume. The control surface offers no resistance to the passage of mass and is arb itrarily located . Analyses to be deve loped later will show that the most suitable location of a control surface in any particular problem is a matter of convenience. Most problems may be treated by either system or control volume analyses. However, the proper choice often leads to substantial simplification. For example, most fluid-flow problems, and especially steady flows, are much more conve niently handled in terms of an appr opria tely specified control volume. Continuum. In the follow ing discussions the fluids und er consideration will be treated as continua. T hat is, their atomic st ru ct ure will be ignored and they will be considered as capable of being subd ivided into infinitesimal pieces of identical structure. In this way it is legitimate to speak of the properties of a continu ul11e.g., density, pressure, velocity- as point p roperties. 2-2
THE LAWS OF MECHANICS AND THERMODYNAMICS FOR FLUID FLOWS
Mass conservation and the continuity equation
The law of mass conservation applied to a system simply requ ires that the mass of the system remain constant regardl ess of its size or shape, the number of pieces into wh ich it is d ivided, or the length of the time interval during wh ich it is observed (for " nonrelativistic" system velocities). Here, as in most cases, the physical law is stated as it applies to a system. This is inconvenient for the stationary observer of flui d flow. Such an observer is unable to continuously measure the behavior of a moving fluid particle (a small system) and must, therefore, content himself with measurements taken at specific points in the space through whi ch the fluid moves . The basic law must then be expressed in terms of the required relat ionships between fluid properties a t various stationary points; Le., at poi nts with in a control volume. Consider a general control vol ume defined within a region of flui d now as shown in F ig. 2- 1. The conse rv ation of mass simply requires that the rate of change of mass stored within the control vol ume be equal to the net inflow of mass: (2-1 ) where m c\, is the mass within the control vol um e and 1i1 is the indicated mass flow rate. These terms can be expressed in more convenient ana lytical fo rms. Consider the mass flow through a small surface element dA on the control surface. If the local velocity is u and a unit vector normal to dA is n (outward poi nting) as in Fig. 2-1, the ele mental mass flow rate is din ~ plul cos adA,
2-2
LAWS OF MECHAN ICS AND THERMODYNAM ICS
9
Fluid flow through a control volume.
FIG . 2- 1.
where p is the local density and a is the angle between uand n, the product lui cos a being positive for outflow and negative for inflow. The mass withi n the control volume me\' can be fo und by the integration lI1,v
~
(
Joy
p d't),
where dU is a volume element and the integral extends thro ugho ut the con trol vol ume. F ina lly, writing the product lui cos a as the usual dot product u . n, the cOllfillllily equatioll may be wri tten
'IIi· Co
j'
P d't)
cv
+ )cs ( pu · n dA ~
0,
(2- 2)
where t he seco nd integral exte nds over the en tire contro l surface. Newto n's second law and the momentum equation
For a system of mass
117
acted on by a force F, Newton's law requires tha t
F
d
~ K -I (mu),
(2- 3)
where K is simply a proportionality constant whose numer ical va lue depends entirely on the units selected for the three variables in the equa ti on. It is possible to choose units of force, mass, length , and time such that K is unity, and this is done in all the following equatio ns. In a derivation very simi lar to the preceding (see, for example, Chapter I of Refere nce [2]), the conseq uences of this system eq uation can be written for a cont ~ o l volume in the following vector equation form, where the notation is tha t of FIg. 2- 1:
L: F ~ ~ 1v pu d't) +
L
pu(u . n) dA.
(2- 4)
The term L:F is the vecto r sum of all rorces acting on the material within the control volume. Forces may be one of two kinds: surface forces (stresses) and
10
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
2-2
body forces . The two types of stresses are pressure, which is a stress normal to the surface, and shear, which is a stress tangential to the surface. Body forces are those which are proportional to the mass or volume of the fluid and include gravitational, electrostatic, and electromagnetic forces. The first law of thermodynamics and the energy equation
The fi rst law of thermodynamics relates heat and work interactions between a system and its environment to changes in the state of the system. To avoid difficulty in application of this law it is necessary to give each of these terms a rather careful definition. The fol lowing definitions are not the only ones which may be used, but they are the basis of the well-established method of expressing the principle of the first law used in this volume. Heat. When two systems at different temperatures are brought together, an interaction occurs. This interaction is ca lled heat. The mathematical sign of the heat is conveniently taken as positive fo r the lowcr-temperature (cold) system and negative for the higher-temperature (hot) system. It is common to spea k of heat transfer, saying that heat transfers from the hotter system to the colder. Thus heat transfer to a system is positive and from a system is negative. If this terminology is used, we must carefully avoid the implication tha t heat is a storable quantity. Heat exists only as a transfer process or interact ion and cannot be stored. A process which does not involve a heat transfer is said to be adiabat ic.
Work. A system is said to do work on its surroundings when the sole effect external to the system could be the rise of a weight in a gravitational field. Work done by a system on its surroundings is, by convention, positive ; work done on a system by its surroundings is negative. Work, like heat, is an interaction between systems, and is not a storable quantity. In general, the magnitUde of the work depends on the motion of the observer. The first law of therm odynamics may be stated fo r a system in the fo rm Q
~
tJ.Eo
+
W,
(2-5)
2-2
LAWS OF MECHANICS AND THERMODYNAMICS
"
both contribute to changes in the internal energy. The second la w, however, shows that they are fundamentally different. It is customary to treat separately those portions of Eo due to motion [kinetic energy m(u' / 2)], position in a gravitational field (potential energy mgz), charge in an electrical field (electrical potential energy qV), and other effects. For most ordinary fluid-mechanics problems it is sufficient to consider kinetic and potential energies only and write, for a system of mass m,
Eo
~
E
+
l/'
m 2:
+
mgz,
(2- 6)
where z is the height above some arbitrary datum and g is the (magnitude of the) loca l acceleration due to grav ity. Further, for cont inua, it is convenient to denote the values of Eo and E per unit mass as eo and e, respectively. Using th is and Eq. (2-6), Eq . (2-3) can be written
Q
~
tJ.
f. (e+ ';' +
gz) p du
+
(2- 7)
W,
J.
where denotes integration throughout the entire system . To express the first law of thermodynamics for a control volume, it is often convenient to introduce heat and work transfer rates. Equation 2-7 may be transformed (see again Chapter I of Reference [2]) in the same way as Eqs. (2-2) and (2-4), with the result that
j
es
2{.dA
~
d/t
(.
j
cv
(eop) du
+j
C8
(h
+ '2" + gz) pu· n dA + (J', -
j
cv
X · u dU, (2- 8)
where
2{. ~ local heat transfer rate (per uni t time per unit area), positive
/0
control volume,
h
~
e
+ pi p
~
enthalpy of fluid ,
net "shear power" (wo rk transfer rate), positive from control volume, and X ~ body force per unit volume on fluid within control volume.
in which tJ.Eo denotes the change in internal energy of a system subjected to heat and work interactions Q and W, respectively.* The internal energy, as its name sugges ts, is a property of the system. The first law in a sense defines this property except for an arb itrary datum va luc. [nternal energy is a storable quantity, in contrast to heat and wor k which may be cons idered energy transfers. The first law itself does not distin guish between hcat and work;
* The notation throughout thi s text differs sli ghtly from the more common fu ndamental texts in that it attempts to present a unified system for the severa l discipl ines covered. Eo (seldom used) refers to in ternal energy inc/lIC/inK kinetic and potential energies, whi le E refers to internal energy without these terms .
The shear power (J'. is that work done by the material wi thin the control volume On the surroundings at con trol surface points where both shear stress and material motion (paraHel to the control surface) occur. Usually this term reduces to shaft p.ower transmitted by shear stresses in rotating shafts. However, there are situatIons where ~he flu.id .may move along a control surface accompanied by a shear stress (at POints WIthin a boundary layer, for example). The last integral, which we shall not often have to cons ider, represents work transfer via body forces . The dot prod uct X . u is the rate (pe r unit volume) at which wo rk is done all the fluid by the environment through body force X (he nce the negative sign precedIng It).
12
2-2
MECHANICS AND TH ERMODYNAMICS OF FLUID FLOW
It is instructive at this point to dwell on certain essential similarities of the continuity, momentum, and energy equations for a control volume. The common form of these equations may be seen by arranging them as follows:
Mass: Momentum:
Energy :
r p du + les r (pu. 0) dA o. d r pu du + r u(pu· 0) dA = L d lev .Ie:; d,
=
.'!.-,
c.t
j.
eop
du +
C\'
r
eo(pu · 0) dA =
) CEI
r 1<. dA les -
components of force and momentum parallel to the axis of the jet, so that the momentum equat ion may be written
lcsr pu·odA.
In each equation the first term represents the rate of change of a certain quanti ty out of the control volume. The right-hand side of each equation may be called a production or source term . The zero on the right-hand side of the firs t equation means, of course, that mass can be neither prod uced nor dest royed. On the other hand, momentum can be brought into being by the application of force. Correspondingly, energy can be changed by heat transfer or work. These equations will generall y be greatly simplified for applica lion in the following chapters, usua lly by assuming steady fl ow and si ngle inlet and outlet states, in which case the integrat ions become trivially easy. I
r---- --T-i Control volume
I
I I
I
L
Fz =
L
Pl/x(u . n) dA ,
in which the subscript x denotes the directio n indicated in Fig. 2-2. The timederivat ive term in Eq. (2-4) has been omitted because the flow is steady. No body forces act in the x-direction, so that the total x-force on the control volume is due to pressure distributed over its surface and the reaction to the
thrust T. We assume the fuel hose to be flexi ble so Ihat it does not transm it force to the control volume. T he force T may be considered to be a normal force on the plane where the control surface inlersecls the strucl ura l me mber which holds the engine stationa ry. The pressure forces always act norma l to the control su rface,
so that only the pressure forces on the two ends of the recla ngular surface of Fig. 2-2 contribute to the x-summation of forces. Further, the pressure on these surfaces may be considered to have the constant value p" except possibly in the exhaust plane of the jet. There it may have a different value p" if the jet is sonic or supersonic. Thus the sum of the forces act ing on the control vol ume in the x-direction can be written as
Ai (::=rL:::;;jl~A~'=f""
~ ;~__~~ !=:::::=;:!-----,- 'T lila
I
In order to show how thrust depends on the properties of the jet, it is convenient to choose the control vo lume indicated in Fig. 2- 2. Its surface intersects the exhaust jet in the plane of the nozzle, where the flow is nearly one-dimensiona l and the momentum flux can be measured more easily than at planes further down stream. The control surface extends far upstream of the engine and fa r from the
with the control volume of Fig. 2-2. In applying Eq. (2- 4) to this control volume, we confine our attention to those
contained within the control volume. The second term is its rate of convection
I
13
ill ustrate the convenience of a suitable choice; but for the moment we proceed
F.
r X.udu lev
_ tP.+
LAWS OF M ECHANICS AND THERMODYNAMICS
floor on which it resls. Later we shall select another control volume, simply to
tIle\'
I
2-2
Illj--""
Reaction to
thrust
in which Ac is the nozzle exhaust area . The momentum flux integra l may be eva luated simply by recognizing tha t the momentum fl ux is infin ites ima l exce pt in the exha ust plane of the jel. There the velocity is norma l to the surface, so that
x
FIG. 2-2.
Iu. nl =
Stationary jet engine.
liz =
li e,
and the momentum fl ux integral is therefore
To illustrate the use of these eq uations in a control-volume analysis, let us consider the generation of thrust by a stationary turbojet. Figure 2- 2 shows the engine running steadily on a test stand. Air enters the inlet at a tota l mass flow rate fila and fuel at a rate rilj ' Since the flow is steady, there ca nnot be a co ntinuous accum ulation of fl uid wit hin the engine, and the rate at w hich the fluid ex hausts from the engine mus t be (lila + tnt). The exhaust velocity is li e, and the force necessary to balance engine thrust is T, as in Fig. 2-2.
L
Pllr(U . 0) dA = /, p,lI; dA, .
,
in whic h Pc is the fl uid dens ity at the exhaust plane. Cons idering the now in this plane to be uniform (i.e., one-dimensional), the integral may be simplified to Pcu; Ae. The mass flow ratc at this plane is PclleA/;l and lhis has already been shown to equal (dIu + 1i1j ). Thus the momentum flux from the engine is
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
14
2-2
THERMODYNAMICS OF GASES
2-3
15
where dS is the change of entropy during a reversible heat exchange dQ at the temperature T. Keenan [I] and other thermodynamicists discuss the concept of thermodynamic reversibility extensively. For our purposes it is adequate to note that processes which involve any of the following are /lot reversible: (I) Friction, (2) Heat transfer with finite temperature gradient, (3) Mass transfer with finite concentration gradient, (4) Unrestrained expansion.
x
FIG. 2-3.
(Ii1.
+
Stationary jet engine.
In general, irreversibility lessens the engineering usefulness of any given process. Though it may be impossible to absolutely eliminate irreversibility from any real process, its effects may often be substantially reduced in practical devices, so that a reversible process is the limit of possible performance. For any process it can be shown that
li11)1I" and the thrust may be obtained from Eq. (2-4) as
'T
~ (1i1.
+ 1;'1)11, + A,(p,
- P.)·
Consider now the alternative control volume III indicated in Fig. 2-3. If we assume the flow in the engine inlet plane to be approxi mately one-dimensional, with pressure Pi and velocity lli. then by reasoning very similar to the foregoing it may be established from control volume III that
'T
~
(ma
+ li11)1I,
-
li1. 1I,
+ A,(p ,
- P.) - A ,(p, - Pal·
This thrust equation is of course not so convenient as the previous one, because it contains three additional terms: Ai. ll i. and Pi. Other choices of control volume could lead to even more inconvenient results. The two thrust equations can easily be shown to be equivalent, however. Using control volume Il and Eq. (2-4) again, it may be shown in a similar way thai
o ~ m.lI, -
A,(p. - p,).
The second law of thermodynamic s and the entropy equation
The second law of thermodynamics states that it is impossible for a system to describe a cyclic process (i.e., to go through a process and return to its original state) which produces work and exchanges heat with a single reservoir of uniform temperature. Although we shall not use this statement directly, we shall use several of its important corollaries. The absolute temperature scale which we shall use is in fact based on this law [I]. Further, the second law permits the definition of the property entropy. For a system,
Q ( -dT ) re v ,
-
(2- 9)
(2- 10)
T
where the equality holds only for reversible processes. Thus if a process is reversible and ad iabatic (dQ ~ 0), it must be isentropic (dS ~ 0), although the converse is not necessarily true. Fi nally, defining s as the entropy per unit mass of a continuum substance, the control-volume form of Eq. (2-10) can be written
f ~TdA cs
:0; -dd t
j.
ps du
c\"
+
f
c~
s(pu· n) dA,
(2- 11 )
where again the equality holds only for reversible processes. 2-3
Using this equation, we can easily transform the second thrust equation to the first one.
dS ~
dS> dQ,
THERMODYNAMICS OF GASES
A pure substance may be defined , in the thermod ynamic sense, as a substance which has only two independent static properties (in the absence of electricity, magnetism, and capillarity). Most engi neering fluids, including gases like oxygen, nitrogen, and vapors in equilibrium with their liquid phase are pure substances. Consider a small system composed of a pure substance in the absence of gravity and motion (of system relative to observer). If the properties are uniform throughout the system , the first law for an ~ncremental change of state is
dq
~
de
+
dlV,
where q and \11 are the heat and work per unit mass, respectively. Suppose the system experiences a reversible process for which dw = p du, in which u is the specific volume. Then, using Eg. (2- 9), we write Tds
~
de
+ pd,.
(2-12)
16
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
2-3
THERMODYNAMICS OF GASES
2-3 6
Although this equation has been derived for a reve rsible process it must also be true for irreversible ones, si nce there are on ly two independent static properties for a pure su bstance, and all the terms in this equation are properties. The entropy change ds depends only on the change in state of the system. wh ich is fixed by the values of de and duo The increment ds is therefore independent of the process required to produce the increments de and du. Th us, with the specified restrictions, Eq. (2-12) must hold true fo r any process whose end states could possibly be connected by a process for which the work is fp duo The solution of a thermodynamic problem in volving pure substances must include sufficient information on the relationships between the properties of the substance. The existence of only two independent static properties greatly simplifies the presentation of this information. If it is given algebraically, th e equations are known as the equations o/state of the substance. Equa ti ons of state are merely statements of experimentally observed relationships. Often the empirical rel ationships cannot be conveniently reduced to eq uat ion fo rm a nd the equations of state must be represented by graphs or ta bles. Eq uatio ns, graphs, or tables are available for the properties of fluids of commo n interest. O ne of the simplest of these is the perfect gas.
Pelfect gas. A perfect gas is here defi ned by the foll owing equation of state: pu
~
RT,
=
R
= T, M
1.9857 calr K ·gm-mo le. Most rea l gases have a large ra nge of temperatures and press ures over which they obey Eq . (2- 14) ve ry closely. By use of the Maxwell eq uat ions it may be shown [I] th at the internal energy of any gas whose state can bc so desc ri bed depends only on temperature: e ~ e(T). Since the enthalpy II is defined by
~
6
~
V
~
h(T).
I COz
II
0, Hz
/
yr::::===
CO
,,-
H
'
0 ,
1000
2000 3000 Temperature, OK
4000
5000
FIG. 2-4. Specific heats of gases [5] . [n this case the relationship between internal energy, enthalpy, and temperature may be described by defin ing the "specific heats" at consta nt volume and tempera ture as follows (for a perfect gas):
de
Cv =
eft'
cp =
dii
dT
(2- 15)
These terms arc ac tually misnomers, since heat is not a storable quantity. Also, the interna l energy and the enthalpy can change in the complete absence of heat. Using the definit ion of enthalpy and Eq. (2- 13), we have
dh ~ de
+ R dT.
The relations (2-15) show that for a perfect gas the specific heats must be related by Cp ~ C,
+ R.
(2- 16)
Typical variations with temperature of the specific heat of gases at constant presSure are shown in Fig. 2-4. Figure 2-5 il lus trates the correspond ing en Lha lpytemperature dependence fo r a number of common gases where the zero has been arbitrarily set at 300oK. Two types of perfect gas calculations may be demonstrated. . 1. Constallt specific heats. Figure 2-4 shows that for a sufficie ntly small range In ~emperat ure it may be acceptable to approximate the specific heats by constants.
USIng the perfect gas law, Eq . (2-12) may be written
+ jJlJ ,
Eq . (2- 13) indicates that the enthalpy also depends on ly on temperature II
/
~
(2- 14)
in which M is the molecular weight and R is the uni versa l gas constant which has the value R 1545.43 ft· lbJio R·lb ,,,-l11ole
h = e
~ 2
(2-13)
in which p is the absolute pressure, v the volume per unit mass, T the absolute temperature and R a constant, called the gas constant. The gas constant depends only on the identity of th e gas . More generally, it is observed that pv
4
8
I
1
~
17
els
dii
ell'
~ T-p't
Or
cis
elT
cI'T -
R
".1:. p
(2- 17)
18
MECHANICS AND THERMODYNAMICS OF FLUID FLO W
2-3
~ 60~--+---~---+----r-~~~~ ~
19
Eq. (2- 17) may be integrated in the form
E
1l
THERMODYNAMICS OF GASES
2. Variable specific ileats. The functions il(T) and e(T) are tabulated for a number of gases (see, for example, References [4) and [8)). By defini ng a function
80 f---+-+-- j--+--',
E
2-3
50 ~--+---+---+--fr-f--+/~
s - s, ~
~ 40
R In 1'_ , /h
(2- 22)
in which the subscript I again denotes a reference value. The function
"~ 30~--+---+-f7f/~<.c
..L
=
e{4J-tP l> / H.
p,
10 I------j~~,p-.'+-+-_t_____j
For any given datum state I this isentropic pressure ratio p, may be expressed by 2000
4000
Temperature, OK
FIG. 2-5.
Enthalpy-temperature relationships for common gases [8].
For the particular case of constant specific heat and for an isentropic process, Eq. (2-17) may be integrated, with the results that
(~t"·
(2-18)
Defining the specific heat ratio 'Y,
and using Eq. (2-16), Eq. (2-18) may be expressed in the form
l!... ~
(I.-)YI<>-".
(2-19)
T,
p,
Using the perfect gas law, this can be transformed to
which may also be tabulated as a function of lemperature. Then if the temperature changes from any value TA to another value Till the isentropic pressure change may be found from
PE
p,(Til)
P.·'
",(T A )
,
in which p,(Tn) and p,(TA ) are obtained from tabulated functions. The state of a gas having only two independent properties may be designated by a si ngle point on a graph whose x- and y-axes correspond to any two properties. A process between states may be indicated by T a line. The temperature-entropy diagram P, shown in Fig. 2-6 is an example of such a graph. . The se ries of lines marked with various pressures correspond to possible constant-pressure processes. The shape of each line may be obtained from Eq. (2-17), L-~
!'-~
(2-20)
!'-
(2-21 )
p,
or
p,
in which p
I/ u is the fluid density.
_________________
s
FIG. 2- 6. Constant-pressure lines on a T-s diagram.
From this it is clear that the lines have pos itive slopes, increasing with temperature. Further, if c1) is a function of T only, the slope of the curve is a function of T only. Hence the curves are of the same shape displaced horizontally.
20
MECHAN ICS AND THERMODYNAM ICS OF FLU ID FLOW
2-4
From Eq. (2-17), it is also clear that
nt higher values of Since this derivati ve is always positive , the upper curves represe pressure. 2-4
EQUILIBRIUM COMBUSTION THERMODYNAMICS; CHEMICAL REACTIONS
ents change Up to this point thermo dynami c processes in which system compon is of vi ta l tion combus Since ussed. disc been not their chemical identity have how chemic al importa nce to many propuls ion method s, it is necessa ry to show is possibl e to sim reaction s are go verned by th ermody nam ic laws. Fortun ately it Fi rst, chemic al plify the discuss ion a great dea l by adoptin g certa in restrict ions. between gases reaction s of interest in propuls ion systems nearly always occur Second , with gas. perfect -a by r behavio in imated which can be closely approx nt chemic al importa the es, process ion expans rapid very th e excepti on of a few in equ ilibrium . processcs usually take place between states which are approx imately gases capable of T he basic problem may be stated as follows: G iven a mixture of will occur, a chemic al reaction and a set of conditi ons u nder which the reaction be asked might one e, exampl For ion? react al chemic that of lt what will be the resu t produc ts of a to determ ine the compos ition and temper ature of the exhaus te method s for constan t-pressu re adiabat ic combus tion process . To formula separat e topics: answering this kind of quest ion, it is convenient to consider three (I) Mixtur es of gases, (2) Chemical transfo rmation , (3) Compo si tion of pioduct s.
Mixtures of gases ly exist as a The reactan ts and produc ts of a chemic al process each general mixture may be gaseous mixtu re of chemic a] compou nds. The propert ies of the Gibbs-Dalton Law, determ ined from the propert ies of the constitu ents by using the wh ich is express ed by Keenan [I ) as follows : pressur es of I. The pressur e of a mixture of gases is equal to the sum of the at the temmixture the of each constitu ent when each occupie s alone the vol ume peratur e of the mixture . respectively, to 2. The interna l energy and the entropy of a mixture are equal, when each ents constitu its of es entropi the and the sums of the internal e nergies m ixture. the of ture a temper the at ure xt mi the of occupie s alone the volume
EQUILIBRIUM COMBUSTION THERMODYNAMICS
2-4
21
Thus for a mixture of 11 constitu ents:
= Tz
...
Tempe ratu re
T m = Tl
Pressure
Pm ='PI +]12 + 1'3
Volume
'U m
Energy
Em
Entropy
Sm
Enthalp y
H"
=
T/!,
(2-23a)
+ ... + PIl,
(2-23 b)
= 9))I Vl = 9R em = 9)1,e, + 9R msm = 9)I , s 1 + IDLJ1m = 9)1 1"1 + 9)}mvm lll
9)l2 V, 9)/,e, 9)I,s,
=
...
=
mCnvl/ ,
+ ... + IDlne, + + BRlIs
"2 +
9)1 2
!,
ll ,
... + 9)l"",,,
(2-23c) (2-23d) (2- 23e) (2-23f)
subscri pts 1,2, in which ID] signifies mass, subscri pt In refers to the mixture and uents. constit of ••• , 11 refer to a series Since de = c, dT, then
+
9)I,c,.1 + 9.1I,c"
(2-24)
Simi larly, C plII
9J"I.o.'C:'/if''iT-'+-·-· _·_+~:,m':.I,!:"C::.t/ll! '. c"I'ccl _+-,--'. 9)-1,,1 = ,, 9)1",
t alone in The pressur es P j, Pz, Pa. Pn that would be exerted by each co nstituen partial the called are mixture the of ature the volume of the mixture at the temper of number the to lated re is e pressur partial pressure s of eac h constit uent. The i, ent constitu any For . present moles of a consti tuent
or, in te rms of the unive rsal c0as consta nt
R, 11.
PiVi=
M~Tm'
where M i is th e mo lecular weigh t of constitu ent i. Since Vi
'U m = ~R i '
Pi =
~n whi~h
9.1] ,.J?'TI/j
-=..--
M /U IIl
Since is the !lumber of moles of constitu ent i present in the mixture . tiai par he t of m su aCCo rding to the Gibbs- Dalton Law, the mixture pressu re is t he pressures, fl i
1~
Pili
IIi
(2- 25)
22
M ECHANICS AND THERMO DYNAMICS OF FLUID FLO W
2- 4
T hus the ratio of the partial pressure to the mi xture press ure is just the ratio of the num ber of moles of the constitue nt to the total number of moles. This ratio is called the mole fraction, X:
(2- 26)
co nst ituents , eac h behav ing as tho ugh the other consti tue nts were not there. It is interes ting to no te that during a process wh ich is isent ropic fo r th e mixture , the co nsti tuents do not necessa ri ly unde rgo isentro pic changes . There is ge nerally a redistribution of e ntro py amo ng the co nstituents.
In determ ining part ial pressures, it is convenient to use moles rather than unit masses. For this purpose one can define molal specific heats by the equations
To avoid confusio n the va lues per mole of a property are usua lly denoted wit h a bar, as
23
In this example, as in all exoth ermic reactions , the hea t o f formatio n is nega tive,
since the system must give off heat to maintain (or return to) the reference temperature. As anot her exa mple, the formation of monatom ic o xygen from nat urall y occurring oxygen is ~02 ~ O. In this endothermic reac ti on the heat o f formation
(+59.162 T he entropy cha nge of the mixture is the sum of the entropy changes of all the
EQ UILIBRIUM COM BUSTION THERMODYNAMICS
2-4
~cal/gm-mole
at 3000 K) is positive, since the heat transfer to the system
must be p()sitive.
These heat quantities may be determ ined experimentally with a simple consta ntpressure contain er. Practically , it may be difficult to produce a reacti on at constant temperatu re and pressure. It is not necessary , in any eve nt, since fr om the stand-
poin t of thermodyna mics only the end states of the reacti o n are of interest. The intermediate conditions do not ma tter (altho ugh they may be important to the preserva ti on of laboratory equipment).
A sleadr jlolV process is another ex peri mental method for determi ni ng the heat of reaction Q/ . Equation (2-8) states that, for a stead y-fl ow process with no shaft work, the heat transfer of the fluid flowing through a control volume is equal to the change in enthal py (neglecting changes in veloci ty and elevation). T hus
dh = Mdh = 1'" dT.
Q = MI or, per moM o f product formed,
Chemical tran sfo rmation
Q A general c hemical reac tion, in w hic h a moles of reac tan t A a nd {3 moles o f reac ta nt B, etc., combi ne to form J.L mo les o f prod uct M plus v mo les o f product N , etc ., may be rep resented by
a A + {3B + . . .
-->
jJ.M + vN
+
(2- 27)
The coefficients a , {3, etc., are kn own fr om the mixture ratio of the reactants, a nd it may be assumed for the prese nt that the prod uct co mposit io n, that is, /.L, v, etc. , is also know n. Accompanyin g a chemical reacti on there is genera lly an excha nge of energy with the surro undin gs . Those reac tions w hich tend to give o ff hea t are ca ll ed exothermic and o nce star ted they will , under the proper cond iti ons, proceed unaided. The other kind , ca lled endothermic, req uire an energy input to proceed. Com bust ion processes, for insta nce, are exotherm ic. T he magn itud e o f these energy inte rac tio ns ca n be determi ned from the compositi o n a nd the heat s of formatio n o f the reac tan ts and products. The co nsta ntpress ure heat of formation of a substance is the heat in teracti on which occ urs whe n one mol e o f the substa nce is form ed at constan t press ure and temperature from its elements as they occur in nature . For examp le, the hea t of for mation of car bon diox ide ( - 94. 052 kca l/ gm -m ole) is the hea t interact io n accompanyin g o the follo wing reactio n from reac tant s to products at 300 K :
=
Q,
=
Hproduct -
H rcactant;;.
(2- 2 8 )
For example, the formation of carbon dioxide in steady flow is represented by Fig. 2-7. Reactant s at T1
Steady-flow reaction chamber
O2
Product at Tf CO 2
t
Qleo, FIG . 2-7 . Schematic of stea dy-fl ow process for formation of CO 2 , with Tj being the reference teJ1lperatu re.
Table 2-1 gives the heats of format ion for some commo n products of combustio n, T/ being the reference temperature at which Q/ is measured. The heats of format ion of natura ll y occurrin g elements are, by defi ni tion, ze ro. ] n co ntr;lSt, most cbmm o n reactio ns are com plex in the sense that the reacta nts are not na tu ra ll y occurring elements and the prod ucts incl ude more tha n o ne substa nc~ . Also, they do not ge nerally occur at co nsta nt temperature. Comple x reac tIOns cj n ~e treated as a co mb inatio n o f simpler reac tions by the followin g meth ?d . (o.nslder the steady-ft o.w r~ac tion ind icated in F ig. 2-8, where Q is the heat 1I1terac lJo n necessa ry to ma1l1ta1l1 constant temperature. It is conv enie nt to consider tris reacti on to be the Slim of dist inct decomp os it ion and form ation reactions.
2-4
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
24
Table 2-1
Table 2- 1
Heals of Formation
Formula
Compound
formation,
Tf , oK
(I)
5 5 5 9
(g) +124.5 (g) + 97.2 + 23.050 (g)
298 298 300
5 5 6
(g)
298 300 300 300
5
Boron Boron atom Bromine atom
B, B
Carbon Carbon (graphite) Carbon dioxide Carbon monoxide
C C CO, CO (CH,,),NNH,
Ammonium nitrate Ammonium perchlorate
Aniline
D imethylhydrazine (unsymmetrical) (UDMH) Fluorine
Br
Fz F
Hydrazine Hydrogen Hydrogen atom
NzH·l H, H HBr HF H202
Hydrogen peroxide
Hydrazine hydrate Hydroxyl
N zH4HzO
OH
Isobutane (2-methylpropane)
C1HIO
JP-3 JP- 4
C7.:10r, H ll.442
Melhyl akohol
CH"OH
Monomethylhydrazine
CH"NHNH2
Nitric acid
* kcal/gm in this casc o
(s)
(g) (g)
Formula
formation,
7 7 10
+ 12.734 (I)
H / C 1.93
HNO"
-
3.0 (I) (g) 0 + 18.906 (g) + 12.000 (I) 1.92 (I) + 52.092 (g) 8.66 (g) - 64.20 (g) - 33.74 (g) - 44.84 (I) - 47.36 (s) - 10.3 (I) + 10.063 (g) - 31.489 (g) - 36. 169 (I) - 26.66 (g) - 0.423* (I) - 48.08 (g) - 57 .04 (I) + 13.109 (I) - 41.35 (I) - 31.92 (g)
300 300 298 300 300 298 298 300
14% N0 2 1% H,O NO N, N NO, N2 0 5 NZ04 N2 0 ::I N,O CSHl S
N itric acid, red fuming N itric oxide Nitrogen N itrogen atom Nitrogen dioxide Nitrogen pentoxide Nitrogen tetroxide Nitrogen trioxide Nitrous oxide N-octane Oxygen
02
Oxygen atom Ozo ne
0 03
-
45.3 21.6 0 +112.5 + 8.091 + 3.6 + 2.30 + 20.0 + 19.55 - 49.88 - 59.795
9 7 5 5 8
300
7
300 300
13
298
5 14
298
10
300
8
8
(g)
(g) (g) (g)
(g) (g) (g)
(g)
SO, H2O
-
57.802 (g)
Complex reaction chamber
300 298 300 300 298 298 300 298 298 300
Reference number
8
5 15 5 5 8 5 5 7
(I)
Water
at T = Tj
12 7
(g)
Sulfur diox ide
Reactants
300 300
(I)
0 (g) + 3.08 (I) + 59.162 (g) + 34.00 (g) (I) + 30.3 - 70.96 (g)
6 10
Tr. OK
kcal/ gm-mole
298
Fluorine atom
Hydrogen bromide Hydrogen fluoride
+171.698 0 - 94.052 - 26.413
(g) (s) (s)
Compound
Reference
298 298 298 300
-399.09 - 87.27 - 69.42 + 6.11
Heats of Formation (Confinued)
number
kca ljgm-mole AL20;; NH,NO" NH,CLO, C(jH5N Hz
25
Heal of
Heat of
Aluminum oxide
EQUILIBRIUM COMBUSTION THERMODYNAMICS
2- 4
300 300 300
7 8
298
5
300
7
~llcts at T = 1j
t
Q Compou nd reactants
at~
Decompos ition reacti on chambers
T
Elemental reactants at T= T..r
Formation reaction chambers
1
Products
Mixing
Multiple products
at T=Tj
chamber
at T = T..r
1
1 F IG. 2- 8. Sc?ematic of complex reaction at constant temperature, showing equivalent e ementary reactions.
26
2-4
MECHANICS AND THERMODYNAMICS OF flUID FLOW
EQUILIBRIUM COMBUSTION THERMODYNAMICS
2-4
The decom position reaction is just the reverse of the formation reacti on ; hence it is a sim ple one. Howeve r hypothetical t he reverse process may be, the first law tells us t hat the decomp osition heat in te raction will be equ al and opposite to the format ion heat interaction. Thus, summing ove r all the decomposition reactions,
Reactants
at T I
(2-29)
Hea l exchanger
Q=HHf - i-i HI
Reactants
Complex
at Tj
reactio n chambe r
t
Q ~ H ltl'!
Products at Tj
Heat exchanger
27
Produc ts at
T2
Q = H p2 - HFf
FIG. 2- 9. Reaction at variable temperatures.
where lli is the number of moles of reactant i and Q/i is its heat of formation per
mole. Once the reactants arc in elemental form , the remainder of the reaction may be treated as the su m of separate form ations, so that (2-30) where Ilj is the number of moles of product j . Since the mi xing process in the t hird chambe r is ad ia batic, t he net heat interaction fo r th e complex reac tion at constant tem pera ture Tj is
L: (n j Q /i)produc ts - L: ( l1 i Q /i)rcactants'
the reference state (Tj ), and may be calcu lated from t ables of the mixture properties or by the methods .of Section 2-3 . The reference tempe rature Tj is simply that temperature for which the heat of react ion, H RP / . is ava ilable. Let us take a typical problem: we wish t o determine the adiabat ic flam e tem perature for a give n reactant mixtu re. Once we know the product composition and det~rm i ne the reference temperature, we can calculate H R b H fl,f, and H RP/. Then, sll1ce Q ~ 0, we can determi ne the product en th alpy H p " a nd fro m this the prod uct tem perature. II
(2- 31)
j
Also, us ing Eq. (2- 28),
Qu =
}fproducls -
I1HI'(
Hrcactnnts =
:L: (njQIi)producta - l:::
I
(n i Q /i) rcnctanb '
Reactants
j
-t--~~~--_t--~T~I---1~----~~~~--------~lf2---+--~T
This difference in ent halpy is given the symbol H ill'! and may be interpreted as the cha nge in enthalpy from products to reactants at th e co nstant reference tem-
/B - - - - - - - - - - - - - - - - -
perature TI . Hence H nPI = H pJ -
H ili =
L
(nj Qfj)prodllct~
-
L
(lI i QI'i )rcuctant~ .
(2- 32)
F,G. 2- 10.
Enthalpy~temperature
diagram for ad ia batic combustion.
j
The va lue of HIl !'! is tab ula ted for many common reactions, but if such a tabulati o n is not available for a particular reaction it can be calculated from Eq. (2-32) and a table suc h as Ta ble 2-\' Note that this calculat ion req ui res knowledge of t he composition of the product. If t he products or reacta nts arc not at the reference temperat ure Tj, the fo regoing procedure must be modi fi ed so mewha t. Hypot hetica ll y, it is possi ble to b rin g the reactants to t he reference temperat ure by heat t ransfer, allow the reaction to occur at consta nt temperature, and then brin g the products to th e final tem perature by anot her heat transfer. The heat interactions required are indicated in Fig. 2-9. The net heat interaction from reactants at T 1 to products at T 2 may be written as the sum of t he three heat-transfer rates indicated in Fig. 2-9. Thus
(2-33)
This ca n a ll be presented graphica ll y on an en t halpy-temperatu re diagram whi ch shows Ihe r elationship between entha lpy a nd tempera ture fo r bo th product and reaClant mIXtu res. Figure 2-10 desc ribes a hypothelical path of the process. The reactants at sta te (Dare brought to T/ a t sta te A, after which conversion to products occ urs a t cons tant te mperature along line A B. Then the prod ucts are heated to t he ~et~t ~et st~te at~ and T 2 is the outlet temperature. Since there is no net heat transinsern.thls partlcul.a: example, ~Rl must t herefore eq ual H p 2 [as may be seen by Ihi tlng Ihe definItion of H ltP! 111 Eq. (2-33) and sell ing Q eq ual to ze ro]. From . s dl3gram It can be see n tha t ir the specific heats* of reactant and product lTIlxtures are unequal , then th e heat of reaction H RI'/ must be a functi on of temperatu re.
---•
Each of the pare nthesized terms is a change in enth alpy from the actual state to
In
>I<
".fhe specific heat of the mixture is give n by the slope of its enlhalpy lem
Fig. 2- 9.
~
l peru ure cur ve
28
2-4
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
2-4
EQUILIBRIUM COMBUSTION THERMODYNAMICS
Composition of products
......
/
]0 all the aforementioned rela tionships, OUf calculations have depended on knowledge of the composition of the products of reaction. Once this composition is known, calculating the state of the product is a rather simple matter. Howeve~, determining the product composition is not always easy. The product composition depends on the temperature, which in turn has just been shown to dep.end on the composi ti on. This suggests that a trial-aod-error procedure may be req UIred for the solution. For certain reactions, it is relatively simple to determine product composition. For example, at sufficiently low temperature, the combustion of a particu lar 110nstoich iometric mixture of octane and oxygen yields the following products:
/
+
150 2 --> SCO z
P.l,/
I Pure A
C,H I s
+
+ bHzO + cO, + dH, + eC + ICO + iO + jOH + kC sH 18 + other hydrocarbons,
150 , --> aCO,
+ gH
(2- 34)
where a, b, c, d, etc., are the unknown numbers of moles of each substance. A reasonable first approximation may be made by recognizing that for all practical purposes Cs l-I I R and the "other hydrocarbon" molecules are present in negligi ?le quantities. However, th ere are still unknowns-a, h, c, d, e, /, g, i, j - for which we can obtain only three equations in terms of numbe rs of atoms. Atom balances yield: For carbon, a e 1= 8;
+ + + 2d + g + j = 18; 2a + b + 2c + I + j + i 2b
For hydrogen, For oxygen,
=
30.
Six more independe nt equations are needed in order to determine th e p roduct composition. They may be obtained from the requirements for chemical equilibrium.
Chemical equilibrium. Consider a reaction in which a stoichiomet ric com bination of A and B reacts to fo rm products M and N with no A and B left: "A
+ f3B --> I"M + vN.
(2- 35)
At the same time thi s reaction is taking place, there is ge nerally some tendency for the reverse reaction, I"M vN --> "A f3B
+
+
A,B,M,N',
I
/
"PJI
"
\
\ Pure M \
P = P.I+Pfj+P.lI+PS
\
I
\
l
T = constant
"
\\
I
II Pure B
Pure N /
jJ/j ~,
I
\
+ 9H zO + 2.50 z.
In this case, the composition is easily determined by requ iring that the number of atoms of each element taking part in th e reaction be constant. Actuatly, at high temperatures this expression may be a conside rable oversi mplification, because the compound s CO 2 , H 2 0 , and O 2 dissociate to some extent into free atoms and radica ls. A m ore general sta tement is
"-
/
/
/
\
C S HI 8
/-- ---- -- - "
29
/
"',........
/
......
FIG. 2- 11.
_------
....-//
Reaction chamber for derivation of equilibrium constants.
to occur, especially at high temperatures. Equilib rium is that state at which bot h forward a nd reverse reactions are occurring at equal ra les. When eq uilibrium obtains, the concentrations of A, B, M , and N are constant. Using the seco nd law of thermodynamics, we can express the conditions for equili brium in algebraic form. Suppose that an equil ibrium mixt ure resulting from the combi nation of A and B is contained in a chamber which communicates through semipermeable membranes with each individual cons tituent substa nce in a pure form, as in Fig. 2- 11. The entire apparatus is surrounded by a constant-temperature bath. A semipermeable membrane is a hypothetical surface through which onc substa nce may pass freely, though all others are preve nted from passi ng. Such membranes are only approxi mated by certain real substances, but they are nevertheless useful ana lytical concepts. For example, in Fig. 2-11, substance A may pass freely into or out of the mixture at cylinder A, and the pressure of th c pure A wilt be equa l to the partial pressure of A in the mixture, whereas B, M, and N cannot pass through this membrane. Similarly, the membranes of cy linders B, M, and N adm it only substances B, M, and N respective ly. The amount of any particular substance within the mixture chamber can be cont rolled by simply pushing or pull in g on the piston in the cylinder containing the pure substa nce. This control process is reversible if done slowly. Starti ng with an equilibrium mixture resulti ng from the combination of A and B in any propo rtion , consider the introduction of additiona l small amounts of A and B through the semipe rmeable membranes into the mixture chamber, in the stoichiometric ratio kaA + k(3B, where k is somc small number. Once in co ntact Within the main cham ber, the A and B tend to react, forming M and N. There are two ways to restore the contcnts of the ma in chamber to their origina l condition. One is to wi thdraw from the chamber kaA + kf3B. The othe r is to with-
30
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
d raw kp. M
2-4
+
- a R Inp., - (3 R Inp B
reached equilibrium, its composition will be restored to its former value. As a result of the reaction
+
k{3 8
->
kp.M
+
~
(d
q) ~
T
k vN ,
":p;''. P PAP H
T
K,- p(CO)p(H ,O) p(CO,)p(H ,)
-
dSnli~tu re = O. ch alllbc r
dS.\(
+kI'S.lI,
dS e
~ k{3SB ,
dS ,y
+ kVSN.
p(NO)p( H 2) [peN))' 'p(H,o)
K _ [p(H,)I'p(O,j r Ip(H 20 )),
In add ition, the partial pressure of each constituent does not change and hence the pre~s ure in each cylinder does not change. Since the temperature is also constant in each cyl inder. the entropy pel' mole of each pure substance must be constant and the change in entropy within the cy linders may be written -kaSA.
K- - ~ )
KG_ l p( H,) )' ' p(O':!) p(H,o)
K _ p(C0 2) ,- p(CO)[p(O, )) ' ,
{3SB - I'Su - vS.y ~ -
~p ~ fi() --r T.
(2- 36)
The entropy of a perfect gas per mole can be obtained by integrating Eq. (2-17) with the result: 7'
-
-
S = -Rlnp+
r
ciT )T\cPT +
p(NO) [piN ,)!' 'lp(O, ))' ,
K - - peN)' I
p eNt)
10 32 10
10'_8
10 - 1
10·2.11 - - -
I
§ 10- Cl
IG
o
.§ 10- '
.~ =s
.t ~
.~
1\
,
10 1
£
1O- !1
"'
~
lOS
\
10.1
10- 13 10
l ;j
:;--"LLL-,-,LL--1_
-L---L---.l
2000 4000 Temperature, oK (a)
-I R npl.
where the subscript 1 refers to an arbitrary datum state. Thus. for a given datum
\
0
10- 11
+
p(CO)
s - p(C)[p(O,)I ' ,
~
or " -SA
K _
[p( H, )J ' 2
o
~ 10- 3
where S is the entropy per mole of pure substance. The total entropy change is equal to thc sum of the changes in the mixture chamber and in the cylinders: k H il I' dS ~ - kaS., - k {3S n + k p.S.l1 + kvSx T
(2- 37)
p( H,o)
.
K,-
dSA
~ K ,,(T).
K, - p( H, )p(O)
For perfect gases it may be shown that H il l' is a funct ion of tempe rature only. Thus for a given , value of k the change in entropy of the system is a function of temperature only : dS ~ kf(T). Since the temperature, volume, mass, and composition of the mixture do not change, there will be no change of state of the system in the mi xture chamber. Thus
~ g(T),
The term Kp is called the equilibrium constant for th is mixture in terms of parLial pressures. Figure 2- 12 gives typical values of the equilibriulll constant Kp. Several general statements should be made regarding equilibrium constants before their use is demonstrated in a detailed example.
kH 1l 1' .
re v
+ p.71lnp.l1 + vRlnp x
whe re g(T) is another!unction of temperature only. Adding the logarithmic terms and wri ting g(T) ~ R In Kp(T) leads to the foll owing relationship between the partial pressures :
there will be a heat interaction of magnitude kH rt p with the constant-temperature bath surrounding the apparatus, since the reaction occurs at the constant temperature T of the bath. In the limit th is heat interaction will be reversible. Then, for a system composed of all the gases in the mixture and in thc cy li nders, dS
31
state, Eq. (2- 36) can be written
k vN. In both cases the total number of atoms of each element in the chamber will have returned to its original value. After the mi xture has again
k~A
EQUILIBRIUM COMBUSTION THERMODYNAMICS
2-4
F,G. 2- 12.
atmospheres.
I?
lO- .,
6000
Equilibrium constants [8] in terms
t)
~K
l, ~
~I
---
I
1000 2000 3000 4000 5000 6000 Temperature, oK (b)
of partial pressures.
Pressure
in
32
MECHANICS AND THERMODYNAMICS OF FLU ID FLOW
2- 4
First, it may be noted that the partial pressures of the prod uct appear in the numerator and those of the reactants in the denominator, each with an exponent taken from Eq. (2-35). Qualitatively, then, one can expect a large Kp for subreactions which approach completion a nd small Kp for those th at do not. The number of constituents considered in this deriva tion is four, but the result can easily be extended to include any number of reactants a nd products. Second, the exponents are the stoichiometric coefficients. For example, consider the relationship between ca rbon d ioxide, carbon monoxide, and oxygen which would be of interes t in our octane~combustion example:
co, -->
CO
+
EQUILIBRIUM COMBUSTION THERMODYNAMICS
2-4
to form x moles of Hand x / 2 moles of O 2 , accord ing to the reaction
xH,O
-->
The relative concentrations of the constituents of the mi xture may be tabulated as shown below. Moles presen t at Mole fraction Xi eq uilibrium Constituents
~02'
p(CO)[P(02)] 1/2 = K p (C0 2 ) p'
H2
x
O2
2:
K II
= XII, ( xo, ) X
where Kp is the equilibrium constant based on partial pressures. Note that Kn, whi le mo re convenient in calculati ng mole fractions, is a function of two variables, pressure and temperature, and is thus inconvenient to tab ulate. In Eq. (2-39), Pm must be in the units for which Kp is tab ula ted. The actual use of these concepts ca n best be demonstrated by an example which contai ns a considerably simplified description of an actua l chemical reaction. ~ 1?,.sec
One mole of H 0 is raised in temperature to 6000°F, and it may be assumed 2 that it dissociates to a n equili brium mixture of H 2 0, H 2• and O 2 at a total pressure of 1 atm according to H 20 ~ H2 + !O 2· The equilibrium constant for this react ion a t 60000 P is Kp = 0.23 atml / 2. The concentration of t he mixture may be determined as follows. During the d issociation, x moles o f H 20 break down
- _ K -
-1/2
pPm
.
Substituting the equilibrium concentrat io ns, we obtain
x
) (
+ x/ 2
x/ 2
I I
1+ (2-39)
1/ '
I[:P
(2- 38)
cr+f3-P.-"K K n - Pm P'
Example 1
1+ x / 2
The relativc conce ntrations may be determined as follows:
(I
Since, according to the G ibbs-Dalton Law, Xi = p;/p,", it may be shown that
1+ x / 2 x 1+ x / 2 X 2( 1 + x / 2)
x
Total
than partial press ures : =
I -x
x
In this case the equilibrium equation is
Kn
X
+ 2: 0,.
xH,
H 2O
It may be noted that these e"ponents a re not directly related to n, c, or f of Eq. (2-34) for the overall rcaction . TI1is equilibrium equation provides one of the six relationships necessary to the determin at ion of product composition after t he reaction (2-34). In a similar way we can state five other equilibrium requirements to govern t he relat ive pa rtia l pressures o f the products. T hird , since the object of this calculation is to determine composition, it is convenient to define an equilibri um constant K n , based on mole fractions rather
33
+ x/2
)
II'
0.23.
x x/2
T;e quant ity x may then be determined by trial and e rror and the mole fracti ons H ,0,0" and H , at equilibrium obtained from the above tabulation TI he foregoi ng calculation oversimplifies t he dissociati on. Strictly spe~king the foI owmg three dlssociat'IOn reactIOns . ' are also of importa nce at 6000°F :
°
O2
-->
20,
H,
-->
2H.
Determination of the equilibrium concentrations of H 0 H 0 0 d H en req Ulre th . I 2, 2, 2, ,an th kno I s e Slmu taneous solution of three eq uilibrium co nditions wit h wn va ues of the three equilibr ium constants for the a bove react ions. Exam ple 2 (from Reference [3])
Consider a complete set of six products for a hydrogen-oxygen reactio n :
AH2
+ B0 2 --> ilH,oH,O + 110,_,0, + I1 H, H ,+ 1100 + I1 HH + /loaOH. (2-40)
34
MECHANICS AN D THERMODYNAMICS OF FLUID FLOW
2-4
Since two equations can be obtained from the conservat ion of hydrogen and
oxygen atoms, four equilibrium relationships are required. The following method is due to Penner [3], who has shown that it is conven ient
srn,
to define yet another equi librium constant based on number of moles rather than mole fract ions. T he use of this equili brium constant results in simplified calculations, and it is introd uced here because the technique has value beyond this example. Using again the typical reaction (2- 35),
EQUILIBRIUM COMBUSTION THERMODYNAM ICS
2-4
35
and the corresponding relationships in terms of moles may be writ ten from Eq. (2-41):
-no - , 1/ 2
st n3
110I-f =
1/ 2
1/ 2
n0 2 1111 2
n0 2
A procedure for obtaini ng a solution for the six un kn owns is now out lined.
STEP I: From the equilibrium equations write the number of moles of each nelV species appearing in the product in te rms of the reactants and the proper equilibrium constant. In terms of sr n :
(2- 41 )
no (2-45)
and usi ng Eq. (2-25),
(2- 42)
nOlI
STEP 2: Use the equilibrium relationships (2- 45) to elimi nate the number of moles
Strictly speaking, the use of this equilibrium constant introduces another unknown, 11, along with an additional equation which simply states that 11 is the total number of moles in the actual product mixture (not the subreacti o n). However, the simplification attained by th is approach lies in the fact that 11 can be rather easi ly approximated, and in the fact that the results of the calculation are not strongly dependent on II. Thus, as illustrated in this example, II is at first assumed known and later checked. An iterative procedure can be used if greater accuracy is desired . Returning to the example in Eq. (2-40), the two equations resulting from conservation of atoms are :
+ 2nH, + nII + nOH ~ 2A, nll ,o + 2no, + no + nOlI ~ 2B,
For hydrogen,
NH ~ 2nH,0
(2-43)
For oxygen,
No ~
(2-44)
where NH and No are the known num bers of hydrogen and oxygen atoms, respectively. The equilibrium information available is of the following form:
of new species in one of the atom-conservat ion stituting Eq. (2-45) in (2-43), we have
-->
STEP 3. Use this expression to eli minate no, from Eqs. (2-45).
no
=
(2-47)
nOH
dln411H2
0, 112
PlI z
Si'''211 ~1,2) . + Jr1l3nl/22
(NH - 211H, 2hl'n411H2
PH ,
Thus, sub-
(2- 46)
"
t02
~quat ions .
STEP 4: Use these expressions along with Eq. (2-46) to eliminate all but the rema ini ng atom conservation equation (2- 44).
1111 ,
from
-
POT-[ 1/2 1/2
jJo zPn 2 pl-r 20
112
PII zPO z
No·
(2- 48)
36
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
Equation (2- 48) can now be solved for
"H. if the various ,\1'" are known.
2- 4
2-4
The
Table 2-2
EQUILIBRIUM COMBUSTION THERMODYNAMICS
37
Enthalpies and Heats of Formation for a Given Product Composition
,fil n's may be eva lu ated from Eg. (2-42) for k~own values of K p , Pm, and n. The
total number of moles in the mixture is actually unknown, but it can be rather easily approximated , and the results are not very sensitive to II anyway. The act ual product compos ition is not Car from that resulLing from "complete" react ion (unless the temperature is very high) so that Il is approximately equal to the number of moles resulting from "complete" reaction. For thi s example, it is
AH2
+
BO, -> A H 2 0
+
2BH,O if there is excess H ,. Hence
Il
+ (A
-
n b.7i
Qj,
kca l/gm-mole
nQj
"
kca ljgm-mole
2 I
0 0
0 0 0
0 0
0 0 0
0.317 0.1008 0.05399 0.1090 0.2333 1.5119
25.693 28.263 16.028 15.897 25.907 36.920
8.145 2.849 0.865 1.733 6.044 55.819 75.455
0 0 59.162 52.092 10.063 -57.802
0 0 3.194 5.678 2.348 -75. 11 8 -63.898
Hz 0, Products
(2B - A)O,
H2
if there is excess O 2 , or
/J.7i,
Reactants
2B)H,
can be readily approximated in terms of the know n
reactant quantities. For a numerica l example. consid er the equi lib rium compos iti on of2H 2 + O 2 at 35000 K (centigrade absolute) and 300 psia. Then NH ~ 4 and No ~ 2. For
0, 0 H OH H, O
complete reaction, the number of moles of products would be 2, and we write 2H2
+
Adding these gives us the tota l number of moles (which was assumed to be 2.2):
O 2 -> 2H,O.
However, since the temperature is rather high, II is assumed to be 2.2. Assumptions of this nature come from exper ience in this type of calcu latio n ; but as we shall see, iterations are easily made and the ass umption is not cr itica L From the
known Kp (pressure in atmosphe res in this case) and pressure (20.42 at m), and the assumed 11, the new equilibrium terms can be calculated . Using Eq. (2-42), ,\1',,1 ~
Kpl C;;,t I12~ 0.17003 , ,\1',,3 ~
K p3 =
1.3046,
,\1',,2 ~ ,\1'" , ~
K P4
Kp, (P~n) -I/2 ~ 0.19362, (~;n)' /2 ~ 15.020.
These values are then used wit h various assumed values of
IlH2
to ca lculate the
left-hand side of Eq. (2- 48). The correct value is of course No ~ 2. T rial results are tabulated below. Calculated No from Eg. (2-48) 2.2561 2.9549 2.0224 1.9917
0.2500 0.3400 0.3000 0.3200
By linear in terpolation between the last two figures,
flH2
was found to be
0.3170. Then Eqs. (2-46) and (2-47) give nH z
Jln
0.3170, 0.1090,
0.1008, 0.2333,
"0
0.05399, 1.5119.
11H z
Il ~
2.326.
The next step is to recalculate "If .. elc., assuming It
~
2.326. When we do this
we find no apprec iable change in -composition. Hence eve n an inexperienced person cou ld probably complete the calculat ion in two, or at mos t three, iterations.
This solution method is not limited to reactions of only two reactants. The generalization of the method for additional reactants is discussed in Reference [3]. In order to take 2 moles of H, and I mole of O 2 from 3000 K and transform them at constant pressure to the above composition at 3500o K, it would be necessary to accompa ny the reacti on by a heat transfer Q which can be determi ned by Eg. (2- 33). If we use Table 2-1 to com pute the heat of reaction we must let TJ ~ 300°K. Since the inlet temperature is also 3000 K in this exa:nple H H - 0 ' III RJ . Table 2-2 shows how the terms (Hp , - HpJ) and H RPJ are com puted from enthalpy-temperature information (e.g. Fig. 2-5), heats of for ma tion (Table 2-2) , and tlelgiven ' .. compositions. 111 the tab le, 11 is th e number of moles and b.h the ent~alpy per mole of eacl~ const ituent above the datum temperature Tj . As before, QJ s the heat of for matIO n of each constituent, per mole. From Table 2-2 the heat of rea c t'Jon H RPj IS . - 63 .8 98 kcaljgm-m ole of oxygen reactant A lso' the enthalpy d'~ (H ' . Th I erence 1' 2 - H PJ ) IS 75.455 kcal/gm-m ole of oxygen reacta nt erefore, from Eq. (2-33), the heat transfer necessary to take two moles of H' and One mole of O at 300 0 K and transform them to the product composit ion o} 2 Table 2- 2 at 3500 K is: 0
Q
(H p2
-
H/2) -
(Hill -
75.455 - 0 - 63.898
~
H HJ)
+
H UPJ
11.6 kcal/gm-mole·02.
38
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
2-4
EQUILIBRIUM COMBUSTION THERMODYNAMICS
2- 4
P .. oblerns
References 1. KEENAN, J. H. , Thermodynamics. New Yo rk: John Wiley & So ns, 1941 H" The Dynamics and Thermodynamics of Compressible Fluid Flow, Volume l. New Yo rk: The Ronald Press Company, 1953; C hapters 1- 8 and 16 3. PENNER , S. 5., Chemistry Problems ill Jet Propulsion. New York: Pergamon Press, 2. S HAPI RO, ASCHER
195 1 4. KEENAN, J. H. and J. KAYE, Gas Tables. New York: John Wi ley & Sons, 1957 5. ROSSINI, F. D., et al. , "Selected Values of Properties of Hydrocarbons." Circula r No. 500, National Bureau of Standards, Washington, D. c., Feb. 1, 1952 6. "Selected Values of Chemical Thermodynam ic Properties." Series III, National Bureau of Standards, Washington, D. c., April I, 1954 7. ROSSINI, F. D., el al. , "Selected Values of Properties of Hydroca rbo ns." Circular No. 461, National Bureau of Standards, Washington, D. c., Nov. 1947 8. MARTINEZ, J. S. and G. W. E L VERUM, JR " Memo randum No. 20- 121, "A Method of Calculating the Performance of Liquid·Pro pellant Systems Containing the Species C, H, 0, N, F, and One Other Halogen, with Tables Required; Thermochemical Proper· tics to 6000o K. " Jet Propu lsion Laboratory, California Institute of Technology, Pasa· dena, Californ ia, Dec. 6, 1955 9. RICE, H. E., "Performance Calculations of RFNA·Aniline and Three·Component Systems Using Hyd rogen." Memorandum No. 9- 2, Jet Propulsion Laboratory, Pasa· dena , California , Ju ne 30, 1947 10. " Heat of Fo rmation of Hydrazine, Un symmetrical Dimethylhyd razi ne, and Monomethylhydrazine." Inorganic Research and Development Dept., FMC Corporation, Princeto n, N. J. , May 14, 1959 11. JOHNSTON, S. A., "A Short·Cut Method for Calculating the Performance of F uels Containing C, H, 0, and N with HN0 3, Oz, or NH-l-N03 ." Progress Report No. 20-202 . Jet Propu lsion Laboratory, Pasadena, California, November 9, 1953 12. CLARK, C. E., "Hydrazine," first ed itio n. Technical Bulletin No. 123.1, Olin Mathieson C hemical Corporation, Balt imore, Maryland, 1953 13. CARTER, J. M., and M. PROTTEAU, "Thermochemical Calculations on the Whi te Fumi ng N itric Acid and JP·3 Propellant Co mbi nation." Research Technical Memora n· dum No. 58. Aerojet.Genera l Engineering Corporation, Azusa, Califo rnia, April S, 1950 14. PERRY, J. H., Chemical Engilleer's Handbook, third edition. New York: McGraw-
39
1. A ?e rfe~t gas of molecular weight M = 20 and specific heat ratio Y = 1.2 expands adia batIcally from a pressure of 1000 psia, temperature 5000 o R, and velocity 1000 fps to a pressure of 14.7 psia. (a) If the fina l temperature is 26000 R a nd the fina l velocity negligible, how much wo rk has been done by the gas during expansion? (b) If.no work i~ done by or on this gas during expa nsion to 14.7 psia, what is the ma ximum pOSSible velocity at the end of the process? (c) If any ad iabatic expansion process to the same final pressu re is permissi ble is it possible for the exit temperat ure to be 300 0 R? State your reasoning. ' 2. An inventor suggests ~ propulsion device shown sc hematically in the figure. Would s~c h a scheme be. feaSible for steady operat ion over lo ng periods of time? D isc uss rigoro usly and bn efly.
Eleclric
I
generalOr
Q
Fluid circuit s
Propel lant
storagc
Nuclea r
tank
~-"--'===--- Cot11ro l ~--_ _ _---.,("""'-~
PROBLEM 2
Rockct Il oule
Hi ll, 1950 15. GAY DON, A. G., Dissociation Energies and Spectra of Diatomic Molecules, second ed ition. London: Chapman and Hall, 1953 PROBLEM 3 3. A strut of cho rd . I d· . except ~ tl '11 CIS P ace In a n Incompressible flow wh ich is everywhere uniform far dOw~\ le I usftrlated effe~t ~f the strut. Ve locity measurements at a certa in plane . . .s ream 0 t lC strut Indica te the presence of a wake as shown T I I·t , . le ve OCI y d IstnbutlOn ac ross the wa k" e IS app rox imately a cos ine function; that is, Il
U
1 -
(J
(1 +
cos 27bry)
for
b -2h < y<2'
40
2- 4
M ECHANICS AN D THERM O DYNAMICS OF FLU ID FLOW
where U is the veloc ity of the un ifo rm flow field fa r fro m the strut and its wake. If the measured values of a and bl e are 0 .10 and 2.0, respectively. determine t he drag coefficient C1) of the strut. The drag coefficient is defined by C1) = TJ/ !pU2 c , where is the total d rag per un it length on the strut d ue to pressure and viscous forces, and p is the fl uid density .
m
Pro pella nt
\
fl ow
,
/h
EQ UILIBRIUM CO M BUSTION THERMODYNAM ICS
2- 4
7, Calculate the combustion temperatu re of an oxygen-hydrogen mixture whose mass ratio is 3: 1 (0 2: H 2). Since this mea ns a co nsiderable excess of hyd rogen ove r the stoic hiometric proportion, assume that the product temperature wi ll be low enough to prevent dissociation. T he temperature of the reactants entering t he cha mber is 20°C, and the heat transfer to the coolant fro m the cha mber walls is 2 kcal per mole of prod uct. As a fi rst approximation, the mean specific heats may be taken as in the table below, an d the heat of reaction as -57.802 kcal/gm-mole at 3000 K (see Table 2-1). Average specific heat, cal/gm·mo le· oC
p"
~
Reactants Products T U RBOJET
R OCKET
P ROBL EM
4
4 . Ca lculate the static t hrusts of the rocket and turbojet engines described in the figure. The th rust T is the fo rce necessary to prevent horizonta l movement of the engine. Both engines ex ha ust a mass flow of 120 Ib/ sec. The ratio of air and fuel mass flowing into the turbojet is 50: 1 and in its exhaust plane the veloci ty is 1500 ft/ sec and the pressure is the same as the ambient pressure . The rocket contains a ll its own propel2 lant and ex hausts it at a velocity of 9000 it/ sec thro ugh an area of 2 ft . The press ure in the exha ust plane of the rocket is 22.5 psia and ambient pressure is 14.7 psia. S. An idea lized supersonic ramjet diffuser consists of an axisymmetric cente r body loca ted in a cy lindrical duct as shown in the fig ure. The flow at stat ionsCD andQ) may be ass umed unifo rm at the values indicated a nd the strea m tube which enters the inlet has a d ia meter D fa r upstream. Show that the aerodynamic drag on the center body and struts is ~ = -
4 D - [Pl tl1 (U2
"
?
-
/1})
+ P2
-
P IJ·
Constituent
6. A mixtu re of gases con taining 10lb of n itrogen, 10 lb o f hydrogen, and 1S lb of helium is contained at a pressure of 100 psia and tempera ture SOOoR . If the constituents are taken to be perfect gases a nd the G ibbs-Da lton law hol ds, what is the molecul ar we ight and specific heat ratio of th e mixtu re?
02
H2O
6.5 7.4
7.5 8.3
11.0
Q" Btu/ lb-mole at 14.7 psia and 7JOF
CHu A
- 130,000
C0 2 (g) O 2 (g)
-1 69,300 0 0 - 104,000
(g) H 2O
5
H,
8. A mixture of air and a hydrocarbon fuel whose average compositio n is indicated by CH5A undergoes complete combustion. Twice as much ai r is present as needed to burn all the fu el. How high is the fi nal temperatu re ? T he heats of fo rmation a re given in t he table be low. It may be assumed that air is 79% nitrogen and 21 % oxygen by volume. T he reactants have a temperature of 77°F befo re combustion . T he adiabatic flame temperature may be fo und eit her by using average values of the specific heats, or enthalpy tables fo r each constituent. A very a pproximate estimate may be made by si mpl y assuming that the product mixture has the same specific heat as air at the same temperature.
N2
PROBLEM
41
9. Find the flame temperature of the products of combustion of hydrogen a nd oxygen whose composition and average specific heats are given in the table below. T he reactants enter the adiabatic combustion cham ber at OaF . In th is ta ble, Cp p is the ~vera~ ~ola r specific heat o f the compo nent between 77°F and t he fla me temperaure, C pr IS the average mola r specific heat of the component between OaF and 77°F ' and Q/ is the heat of forma tio n at 77°F . Component
Gas
Mol wt, M
Speci fic heat ratio, 'Y
02 H2 0
Nitrogen H ydrogen Helium
28 2 4
1.4 1.4 1.67
OR
H
H 2O
Moles 0 .1008 0 .3 170 0 .054 0.109 0 .233 1.512
Cp p,
Btu/ lb-m ole· oR 8.8 8.0 5.0 5.0 8.0 11.2
Cpr Btu/ Q" lb-mole 7.0 7.0
0 0 +1 05,400 + 92,910 + 18,000 - 104,000
MECHANICS AND THERMODYNAMICS OF FLUID FLOW
42
2-4
10. One mole of C02 is heated at atmospheric pressure to 50aaoF. What is the equilibrium mixture composition if only CO, 02. and C02 are present? 11. An inventor claims that a secret device shown figuratively below can take airstreams anctQ) and convert them into output, streamsG) and.c±> without energy exchange with its environment. [nside the device are mechanisms for mass, momentum, and energy exchange between the streams. Is it possible that such a device can be made to operate in the steady state? It may be assumed that ve locities at all stations are negligib le.
CD
3 Steady One-Dimensional Flo"" of a Perfect Gas
PR013LE1vl
11
12. Show that if any cons tant pressure line p J is given on a temperature-entropy plane, a line corresponding to any other constant pressure p can be obtained by horizo nta l shift of magn itude (for a perfect gas):
I1S ~ -R In 1'.. . PI
13. Two st reams of ai r mix in a constant-area mixing tube. The primary st rea m enters with a veloc ity of 1000 [LIsee and a temperature of the mixing tube at station ISOooR. The secondary stream enters with velocity 100 ft/ sec and temperature 50QoR. The flow at stationsQ) andQ) may be assumed one-d imensional. The p~es. sure at s(ationCDis 15 psia and the rat io or prima ry to secondary flow areas at station
CD
(Dis 1:3.
;;'l~::===t-
Primary
""
Sccond--~ar~Y~:..-j-I----;=,II.'_ _ _ __ _i7::'\
CD
PROBLEM
CD
INTRODUCTION
3-1
In most of the propulsion methods discussed in this volume, the working fluid can be quite satisfactori ly app roximated by a perfect gas. In this chapter the basic laws stated in the first chapter are applied to perfect gases flowing in channels . For algebra ic cOllvenien~e the specific heat of the gas is usually assumed constant herein.
Useful analyses of actual channel flows are re latively easy if it is assumed that ~he ~uid co ndition~ vary in t he streamline direction only. In this case, the flow IS saId to be one-dImensional. The flow in t he immediate v ici nity of a waH can never be one-d im ensional, s ince the velocity at the wa ll surface is zero and there are significan t property va ri ations across the s tream lines . However , with the ex~eption of a thin layer of fluid next to the wall , fluid condit ions are often fairly umform over a large part of a flow. Even if the bulk of the flow is not uniform on any s urface norma l to the stream lines , the one·ciime ns ional approx imation may still lead to useful solutions for the st reamwise va ri ation of averacre fluid . 0 propertIes. One-dimensional so lut ions exhibit the major c haracter istics of many Impor tant flows, including t hose in nozzles, diffusers, and combustors.
13
(a) Using co ntinuity, momentum, and energy equations, a l ~l~g with the perfect-gas law, show how the flow at Q) may be determined from condltl~ns atG). (b) Determine the velocity, temperature, and pressure at slalIon(~).
3-2
GENERAL ONE-DIMENSIONAL FLOW OF A PERFECT GAS
Fluid flows arc governed by the equations of conti n u ity, momentum and energy alan . h . ' , WIt · an appropriate eq uation of state. In this sec ti on these equations are applied to a one-dimensional diITeren tial cont r ol volume within a duct , as indi-
?
43
3- 2
STEADY ONE· DIMENSIONAL FLOW OF A PERFECT GAS
44
cated in Fig. 3- 1. The ra tes of heat and work tran sfe r are dQ and dO'" respectively. A body force of magnit ude X per unit volume of fluid may act in the stream direction. For steady flow t he continuity requirement [from Eq. (2- 2)] is Lpu, n dA = O.
Since the flow of Fig. 3- 1 is ass um ed one-dime nsi onal, t his expression may be reduced to (d/ dx)(pllA) = 0, o r
+
+
dp dA dll = 0 pAll
(3- 1)
for the cont rol vol um e of infinitesimal le ngth ind ica ted in Fig. 3-1 .
A
" -
p
r
du II+-:r:..dx
I I
I p+ dp dx(, x
dp
~ ~/.1
TO \
x----
3- 1.
dQ
=
j' Q.
dA"
=
l'iiI
L . ,
(h
u,
?Q;;;;;;:~
Control
dfr
+
d6'"
u du = 0,
F = Lpu(u' n)dA.
+ roCdX) +
XA dx
_(0: + TI)~) + X dx A To
+
Which may be integrated to
Writing 'LF as the sum of ..pressure. wall shear, and body forces, and aga in sidering th e fl ow of F ig. 3-1, this equation can be written
where
d1)
The stagnatioll state is defined as that state which would be reached by a fluid if it were brought to rest reve rsibly, adiabatically, and with out work. The energy equation for such a decelcration is obtained from Eq . (3-3 b),
volume
Iro '" fr
or
r X· u
){'\.
where A{,!l indicates the ent ire cont rol surface area (as opposed to the area A of the duct cross section). The gravitational potential energy term has been neglected, as is usually permissible. Further, except for specia l cases to be co nsidered in Chapters 5 and 16, the body force X will be ze ro. In this case the e nergy equatio n reduces to Ih(dh + II dll) = dQ - dO', (3- 3a) or dfr + II dll = dq - dll', (3-3 b)
The momcntum eq ua ti on for steady fl ow [from Eq. (2- 4)] is
(A dp
pu . n dA _
Stagnation state
>-
One-dimensional flow throu gh a control volume.
L
r (Ir + 1~2)-
)es
]f the fl ow a rea, heat tra nsfer, work, shea r stress, and body force are known, changes in the sta te of lh c fluid stream may be obtained from th ese four equat ions. ]n genera l, solutions cannot be expressed in simple algebraic form. There are a numbe r of important special cases, however, which can be in tegrated in closed fo rm . Before we proceed to these special cases, let us int rod uce two concepts useful fo r all cases: the stag nat ion state and the Mach num ber.
X
I -:-_---~
"wc:::e:r FIG.
For steady fl ow the ene rgy req ui reme nt [from Eq. (2-8)] is
dA
A+Tdx (X
I
1
p
45
where liz is the mass flow rat e puA (a constant) and q and 1\1 a re the heat and work tran sfer per unit mass. If the fluid ca n be assumed a perfect gas (as it will be in this chapter), the equa· ti on of state is p = pRT. (3-4)
d (J's
%
GENERA L ONE· DIMENSIONAL FLOW OF A PERFECT GAS
3-2
=
=
is the wall shear stress and c is th e duct circ um ference.
2
(3- 5)
COIl -
plIA dll
plI!!'!. ' dx
/1
+ "2 .
(3 - 2)
The constant of in tegratio n, fro, is cal led the stagnat ion e nlhalpy . For any ad iabat ic fl ow viewed from a reference frame in which no wor k occ urs the stagna tion enthalpy is co nstant. Note parti cularly that this co nclusion h old~ for both reve rsible and irreversible flows. U nlik e ent ha lpy, the press ure reached When there is ad iabatic zero-wor k deceleration of a fluid is decrea sed by irre~ersible ~rocesses . This is ill~s trated in Fig. 3-2, an enthalpy-en tr o~ diagram Or a typIcal gas. The stagnation state corresponding to state is ® a nd the stagnation press ure is POI. which is the pressure atta ined by the flui d afte r it has
CD
46
STEADY ONE-DIM ENSIONAL FLO W OF A PERFECT GAS
3-2
3-3
ISENTROPIC FLOW
47
where a is the local speed of sound in the fluid. The speed of sound is the speed of propagatIOn of very small pressure disturbances. For a perfect gas it is given by [ I]
IIi
flu]
(3- 8)
2
3-3 pz
p,
M.any a~ tual processes, such as flows in nozzles and diffusers, are ideally isentrop'c. It 15, therefore, worthwhile to study the isentropic flow of a perfect gas in the absence of work and body forces . The simple results obtained with the constan t specific heat ~pproximation are useful even for large temperature changes, so long as approprIate average va lues of cp and)' are used . For constant specific heat, Eq. (3-5) may be written
Entropy
FIG. 3-2.
D efinition of stagna tion state.
been stop ped isentropically. If the flow is brought to rcst irreversibly, as at @ (l he increased ent ropy is a measure of the irreversibility), the stagnation enthalpy is the same as before but the pressure attained is less than Po l' For any process, slIch as to..0), the stagnat ion state gene rall y undergoes a change li ke the one from @ to (Q;). The decrease in pressure from Po 1 to P02 is, in the absence of work, an indication of the irreversibility of the process. If the flu id is a perfect gas, the stagnat ion tel11peralUre is related to the stagnation enthalpy through the
CD
re lation
'I'
ho - h =
ISENTROPIC FLOW
f OcI' dT.
!.-~ = l+!/ T
(3- 9)
2cI'T
which, with the definition of Mach number, becomes To = I
T
+ ~M' 2'
(3- 10)
Equation (3-6) may then be wri tten
7'
The stagnation temperature To of the fluid is that temperature which would be reached upon adiabatic, zero-work deceleration. For a perfect gas with constant specific heats the stagnation temperature and pressure are related [using Eq . (2-19)] by
l Po _ ( 'Y - 1 2)I' /n - ) - I +-- M .
P
2
From the perfect gas la w and Eqs. (3- 10) and (3- 11),
(3- 6)
Po
-=
P
since the deceleration process is, by definition, isentropic. Other properties of the fluid at the stagnation state may be determined from Po, To . and the equation of state . The stagnation conditions may be regarded as local fluid properties. Aside frolll analytical convenience, the definition of the stagnation state is useful experimentally since To and Po are relat ively easily measured. It is usually much more convenien t to measure the stagnation tempera ture To than the temperature T. Mach number
A very conven ient variable in compressible-flow problems is the Mach number M, defined as (3- 7) M = !!.. a
(3- 11 )
(I + )' -- 1 M-")
2
1/("(- 1)
.
(3- 12)
For allY flow of a perfect gas these can be used to relate the local conditions (T,p~ p) t? the local stagna tion conditions To, Po , Po, since this relationship is by definff /oll IsentropIC whatever the actual flow considered may be. The mass flow per unit area is liz A = pli.
Using Eqs. (3-7), (3-8), and (3-10), we may express the velocity as !I
= M
J--
0fFo- -
I
+
-y ;
I M' .
48
3-3
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
10
f II
f~
-
,,-
~-
I2
"-
"-
/
""
I" 'I, I~~ I i
I
A
A*
I I
\
1\ ~ \ .1 \
0.02
I
0.1
0.2
2
0.4
l Po
\
\
4
Mach number, M
FIG. 3-3.
2)J'H ')/2(7-1) .
(3- 15)
1"0 3-4
,
One-dimensional isentropic flow of a perfect gas.
FRICTIONLESS CONSTANT-AREA FLOW WITH STAGNATION TEMPERATURE CHANGE
Anot her special case for which the basic equations may be solved in closed form
I
I
! I
0.0 I
r'
-
a
is frictionless flow in a constant -area duct in which stagnation enthalpy change occurs. Four processes of interest in which the stagnation temperature of a moving stream changes are:
\ I p\:-r" - I
Po
+
perfect gas of y ~ 1.4. Tabulations for other Y are given by Shapiro [I], Keenan and Kaye [2], and others.
_\
.1
,
(3- 14)
appropr iate 'Y. F igure 3-3 gives the one-d imensiona l isentropic function s for a
,
0.04
1' +1
For a given isentropic flow (given 'Y, R, Po, To , 111), it is clear that A* is a constant so tha t it may be used as in Eq . (3- 15) to normalize the actual flow area A. Eq uations (3-10), (3-1 1), and (3- 12) relate the fl uid properties to the Mach number M and Eq. (3- 15) shows how the Mach number depends on the flow area. Expressing fl ow variables in terms of Mach number is a matter of great convenience, since the various flow functions can be plotted or tabulated as a funct ion of M for the
I
I
.
) (1' + 1) / 2(1' - 1)
I [ -2- ( I+ y -I M M y I 2
- ~ -
\\
v RTo
(2 - -
Combining Eqs. (3-13) and (3- 14), we have
"
\\
I
0.1
±
'-
I
A*
I
49
1 with an asterisk, the maximum flow per unit area is, from th Po - ~ -- ...n
V
I\"
I
I
0.2
'
I
,
/
~
I "
I
I
0.4
--
,
,/
/
CONSTANT-AREA FLOW, STAGNATION TEMPERATURE CHANGE
of the flow at M Eq. (3-13),
.f--L-
fA I A'
I-
4
3-4
,
I. Combustion, 2. Evaporation or condensation of liquid drops traveling with stream,
I
3. Flux of electric current through a fluid of finite conductivity (Joule heating), and 4. Wall heat Iransfer.
\0 l' = 1.4
(After Shapiro [11 ·)
Then, using this expression with Eq. (3- 12), we may write the mass flow rate as (3-13)
For a given fluid (y, R) and inlet state (Po , T o), it may readily be shown that the mass fl ow per Unt't area ,'s maximum at M = I . If we indicate those properties
In any real flow, frictional effects are always present, especially near solid boundaries. As we shall discuss in Chapter 4, wall heat transfer and wall friction are in fact so closely related that it is not rea list ic to discuss the forme r with out at the same time talk ing about the latter. Nevertheless, the study of frictionless flow in a consta nt-area duct in which a To c hange occurs illustra tes some important features of the real flows in which the fi rst three effects above are present. Combustion and evaporat ion or condensation processes may result in variations of mo lecu lar we ight an d gas consta nts, but for th e sake of simplicity these may be considered negl igib le. The stagnation-entha lpy cha nge in the processes can be determined fr om I:1h o = q - w,
/
L
3-4
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
50
. For th is case d 1\I are net heat and work transfers . per unit fmass Eof fluid where q an (3 I) (3 2) and . ·t momentum and energy equations are rom qs. the contlOUI y, ' (3-3), respectively: 1
+
dp
Momentum:
dp
Energy:
dll o = dh
-pu dll,
+
II
and, using the above relation between static pressure rat io and Mach number, T r--;
To
dll.
T Ol
The first two of these can be readily integrated to
P _ p,
in which the subscript also be expressed as
pll = PIll!! = - (pII)(U - Ill) = -(P,II,)(II - II,),
sign ifies an initial condition. The second of these may ,
P - PI = PI III - pll
2
. duc·mg the Mach number , we can transform this to By mtro
p _ I + 1'M ; . I + 1' M'
p; Further, by use 0 f Eq.
(3 -11) we can derive the stagnation pressure ratio from ,
this equation, with the result that
(
l' _
I + -
.fJll.. POI
Th
I + 1'Mi) ( I + "'1 M 2
I
,)''''-1)
- M(3- 16)
2 I +
l' ;
1 M;
the change in stagnation pressure is directly related to the Mach number.
;~e dependence of Mach number on stagnation enthalpy (or stagna l10n tem-
perature) can in turn be obtained as follows : Using the perfect-gas law, T Tl
I' P,
=
P-; p '
or invo king the continuity condition, pll = PIll " it may be seen that T
r-; Now if th e Mach num ber reI at shown that
·lon M = II/V"'IRT is introduced , it can e;1sily be T
51
=
[I ++ 'Y M' (M)J2 'Y M ;
I
M~
.
From Eq. (3- 10) and this expression, we can show that the stagnati on temperature ra tio is governed by
II
=
CONSTANT-AREA FLO W, STAGNATION TEMPERATURE CHANGE
,
du = 0,
Continuity:
P
-
3-4
1 +1'M;
[I
=
M - 2 1+ 1'- 2- 1 M-' ) + 1'M' (M)] ( I + M; .
~~
(3- 17)
Thus the Mach number depends uniquely on the rali o of stagnati on temperatures and the ini tial Mach number. It is desirable in this case, as in the isentropic case, to simpl ify these relationships by the choice of a convenient reference state. Since stagnation conditions are not. co nsta nt, the stagnation state is not suitable for this purpose. However, the state corresponding to unity Mac h number is suita ble si nce, as Eqs. (3-16) and (3-17) show, condit ions the re are constant fo r a given flow (i.e. , for a given Po" To" M I). Again using an asterisk to signify properties a t M = I , it can be shown from the equations above that
T T* To T~ ..P....
(1 +1')' , I + 1'M " M- , 2(1' + I)M 2 (I +
l' ;
1 M2)
(3- 21 )
( I + 1'M" F
1+ 1'
P*
I + 1'M2
Po
(_ 2 l' + I
Po*
(3-20)
rn-I) (
(3- 22) I + l' ) (
I + 1'M"
'Y _ I ?)'Y !('Y - 1 ) M. + 2I
(3-23)
In this way the fluid properties may be presented as a fu nction of a single argument, the local Mach number. Equations (3- 20) through (3-23) are shown grap hically in Fig. 3-4 for the case l' = 1.4. G iven particular entrance conditions, To 1, Po 1, M 1, the exit conditions after a given change in stagnation temperature may be obtained as follows: The value M, fixes the val ue of T orlT I; and th~s the va lue TJ, since T O l is know n. The exit state is then fixed by T02 /TJ, detenmned by or Then M "p z/p* ,Poz/p~ are all fixed by the value of T oz/TI; (for a given value of 1').
• 52
3-4
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
3-5 4.00
/
/
I
, ,::::,
l-
0.60
t-
~,
-
\
Ij 0.40
Momentum:
T
\ I '
~
+ u du o
~ dh
For flow in long pipes it will be shown (Fig. 4-14) that the wall shear stresses can be correlated as follows:
p'
//
°
Energy:
\
l/
0. 10 0.08
Continuity:
To
kTZ
/
0.20
CONSTANT-AREA FLOW WITH FRICTION
respectively,
p~ /
- - - - - - 1- - -l-
f1.00 0.80
/
p"
1--
53
Another solution of Eqs. (3- 1) through (3- 4) can be obtained for constant-area adiabatic flow with friction but no body force or work. In this case, from Eqs. (3-1 ), (3- 2), and (3-3) the continuity, momen tum , and energy relations are,
/
I
2.00
CONSTANT-AREA FLOW WITH FRICTION
3-5
TO
or
0.0 6 0.04
0.0 2
0.0
,
diameter.
-
For Mach numbers greater than unity, the Mach number is also a
significant variable. The elTects of roughness (or at least of variations in roughness) are usually fairly un important for surfaces which have a reasonably smooth finish.
II); 0.04
FIG. 3-4.
,
in which C, is the skin friction coefficient To / (pu z/ 2), Re is the Reynolds number puD/ !" ~ is the average height of surface roughness elements, and D is the duct
, 0.1
0.2 0.4 0.6 0.8 Mach number, M
2
3
"( = 1.4
This point is discussed in more detail in Chapter 4. T he way in which friction alTects the Mach number may be shown as follows. First the conservation equations are transformed (by ihtroducing the perfect gas law and the definition of Mach number) to read:
Frictionless flow of a perfect gas in a constant-area duct with stagna tion
It may be seen from Fig. 3-4 that inc reasing the stagnat ion tempera ture drives the Mach number toward un ity whether the flow is supersonic o r subsonic. After the M ac h numbe r has approached unity in a give n duct, further increase in stagnation enthalpy is possible only if the initi al conditions cha nge. The flow may be
said to be thermally choked . It is int eres tin g to note the variations in sta gna ti o n press ure as the stream is subjected to energy transfer. The stagnation pressure alwa ys drops when energy is added to the stream and rises when energy is transferred from the stream. Thus, whether the flow is subsonic o r supersonic. there may be s ignificant JOSS of stagnation press ure due to combustion in a moving s tream. Conversely, cooling tends to increa se the stagnat ion pressure.
Continuity:
temperature change. (Afte r Shapiro [1].)
T
2 u-z.
2
0,
= " M 2 (4 C/ dX) _ 2 D
Momentum:
Energy :
~ dll
o ~ dT
T
+
,,- l lIdll RT'
- ,,-
Since
M
the n dM 2
M2
II
V"Y RT '
dll , RT
II
54
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
3-5
3-5
CONSTANT- AREA FLOW W ITH FRICTION
55
/00
And now, if we use the con tinuity eq uation to eliminate du 2 / U 2 , it follows that
60 40 20
Also, combin ing the momentum and energy eq uations, we can easily show that
\
\
Po
~, I
dp = - 1'M' (4 C dX) P 2 D j
+
0
_1'_ dT 'Y- I T
6 4
Combining these last two expressions yields
dM' = M'
-M ' Y
(4C dX) D
2
,
,, ~
1/
f
I ; /
1
T T'
which, for constant stagnatio n temperat u re [us ing Eq. (3- 10)], becomest 0.6
4C U
0.4
/ DII
(3- 24) 0.2
Equation (3- 24) shows that dM' > 0 for M < I and dM' < 0 for M > I. That is, frict ion a/ways changes the Mach number toward unity. For constant Cj, Eq . (3- 24) ca n be in tegrated between the limi ts M = M and M I to yield
V\ 1\
/ \
I
0.1
~\
P'
0.06 U 4Cf D
0.04
(3- 25) 0.02
in which L* is the le ngth of d uct necessary to cha nge t he Mach n umbcr of the flow from M to un ity. Co nsider a duct of given cross-sectional area and variable length. If the inle t stale, mass flow rate, and average skin fr iction coefficient are fixed, there is a maximum length of duct which can transm it the flow. Since the Mach number is unity at the ex haust plane in that case, the length is des ignated L* and the flow may be said to be fric ti on-choked. F rom Eq . (3-25), we can see that at any poin t I in the duct, the variable Cf L* / D de pends o nl y on M 1 and 1'. Since D is constant and C, is assumed cons tant , then at some other poin t a distance x (x < L*) downstrea m from point I ,
0.01 O.t
0.2
0.4 2 Mach number, M
4
/0 1 = 1.4
FIG. 3-5. Ad iabatic flow of a perfect gas in a consta nt-area duct with friction. (After Shapiro [\].) by lIsing the definition of Mach number and the continu ity and energy re lations are : T
1' + 1
(3- 26)
T*
p
•
(3- 27)
From this we can determine M z . Other relationships which may be derived [I] Po
t Fo r noncirc ular ducts, the hydraulic diameter mation in place of the diameter D .
DII
=
4A / c can be used as an approxi-
Po*
(3-28)
56
3-5
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
1.7
1.5
I-
0.9 1'1
1)::,..
!,.?
_}-I~
1-p<, l:V'l' /~ 0' Pj ,"-.Vi
'" I
o. 7 L-~~~~~~~~~~~~~~~~~ o 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.t8 s
FIG. 3- 6.
Typical temperature-entropy diagram for choking process (subsonic).
SHOCKS
3-6
57
Equations (3- 26), (3-27), and (3-28) are plotted in Fig. 3-5 for l' = 1.4. Again the asterisk denotes properties at that point where M = 1. In Sections 3-3, 3-4, and 3-5, three special problems have been treated to show the separate effects of area change, energy transfer, and fr iction. In each case, reference was made to a state where M = 1 and properties at that point were disti nguished by an asterisk. Note that properties at that point are used as convenient normalizing quantities only, and the state need not actually exist in the real flow studied. For a given inlet condition, the M = 1 conditions represent three distinctly different states. This is clearly indicated in Figs. 3- 6 and 3- 7. For two initial states, one subsonic and the other supersonic, three processes to reach a Mach number of one are shown on a typical T-s diagram. The energy-transfer path is called a Rayleigh line and th e friction path is called a Fanno line. On both these lines the * condition occurs on the nose of the curve. The notation *s, *J, *11 indicates the M = 1 states reached by isentropic, friction, and heating processes, respec tively. Subsonic states are above the nose of the appropria te curve, while supersonic states are below it. Note also that travel in eit her direction is possible along the isentropic and energy-transfer curves, while only one direction is possible on the frict ion curve. If it is important to study combined effects of these processes, then simple solutions are not available. However, the complete equations may be integrated numerically according to procedures described by Shapiro [I].
3-6
SHOCKS
A shock is a discontinuity in a (partly) supersonic flow fluid. Fluid crossing a stati onary shock front rises suddenly and irreversibly in pressure and decreases in veloci ty. It also changes its direction except when passing through a shock which is perpendicular (normal) to the approaching flow direction. Such plane normal shocks are the easiest to analyze.
'L T01
1.0
Norma l shocks
F or flow through a normal shock, with no direction change, area change, or work done, the continuity, momentum, and energy equations are
,
I
0.10
0.12
0.t4
0.16
s
F IG. 3-7. Typica l temperature-entropy diagram for choking process (supersonic).
Continuity:
PIli}
Momentum:
PI - pz =
Energy :
To 1 = T 02 ,
=
P2 U 2 , PIUl(U2
-
lit),
where the subscripts 1 and 2 indicate initial and final states, respectively. Two solutions to these equations are possible, one of which states simply that no change
58
STEADY ONE-DIM ENSIONA L FLOW O F A PERFECT GAS
3-6
3-6
SHOCKS
59
occ urs. In term s of Mach number, the interes tin g solu tion may be stated [I ] th at Co nt rol volume
M' + _ 2_ 'Y- I
I
I
M~
., 2'Y - - - M; 'Y-I
,
! 'Y 'Y
+
I L _ _ _ _
I I '
)]Y!(Y-l) + - -- M; 'Y -
.
III
Streamlines
s
1
2
P02
s
PO I
FIG. 3- 9.
T?
Oblique shocks
T; = In F ig. 3-8, M , and PO,/POI are plo tted for the case 'Y = 1.4. Note especially th a t rather large losses of stagn ation press ure occ ur across Mach numbers greater than abo ut 1.5. Tables of all the functions for var io us 'Y are provided by Keenan and Kaye [2]. I. I'\.
o
o. 8
o. 6
"
""
I
""-
P02 =
'\.
1\01
l iT
I
SlalgnatIO[~ pre~sur~ I 1,1
1.1;10 I-
"'- "- ~ M,l---'\
o.2
o. t F IG. 3-8.
If a plane shock is inclined at an a ngle to t he fl o w, t he fluid passing through suffers n ot only a sudden rise in pressure a nd dec rease in speed but a lso a s udde n change of d irec ti on. The situat io n is illustrated in F ig. 3- 9 fo r an obliq ue shock s-s' in one-dimens ional flow. I n passing through the shoc k t he fluid is deflected through a n an gle O. The basic equati ons ap plied to the indicated control volume are : Continuity:
Pllfl l l
Energy:
TOl
1
P2 1l 2n,
T 02 .
r
Downslrca: 1
\
I
jJ I -
(b) Momen tum parallel to the shock,
o=
IIU
'Y =
1.4
1\ 4 6 2 Upstream Mach number, M I
8
.,
(a) M ome ntum normal to the shoc k,
jJ2
=
?
P 21f21l -
Plill ll (1I21 -
111
P lili n.
1111),
in wh ich the subscripts f1 and ( indicate d irect ions normal to a nd parallel to t he shock, res pect ively. F rom (b),
1\
\
=
=
Two momentum equations may be written for this flow, o ne for changes mo mentum perpendicu lar to the shock, the olhe r for cha nges pa rallel to it :
Mach n umber
o. 4
Oblique shock geometry.
10
Normal shock fun ct ions. (After Shapiro [1].)
=
Ill!'
Since the veloc ity co m ponent pa ralle l to t he shock is the same o n both sides of it. it may be see n that an obl iq ue shock becomes a normal shock relative to a Coo rd inate systcm moving with ve locity "11 = " 2 /. This fact may permit us to Use the normal shoc k equations to calculate obl ique shocks. For cxample, o-ive n the. upstream (or initial) Mach number M J a nd th e overall stagnation pre~sure ratIO, the compone nts M In and M 21< could be determined from Fig. 3- 8. The presSU r e a,nd t cmp~ra ture ratios p 1 a nd T 2/ T I co uld th en be obtained by rep lacing M 1 With M III In the appro priate normal shock eq uations. T he shock angle 0' could
ez/
D efl ectLOn ang 1e 0 - 0° ( n ormal shock)
90
\ 80 \
\
\
II "
7
/'
---
1- -
V
I
/
'l_
o \
I
-t--
l.--~V:::--l-/ ./
/
-+- --{-=- M''''''[
\
SHOCKS 3.>-
---
1
f
--
35~
(min)\
Of
o
3-6
I
2
V IL
3.0
~ t'--- ---l\~ '"\ ~'" ~ "~ r----- '?i--- -
~
-
i
MaChD7IV !/
t-
O
~~~ = ~ t-t-w~~~ ""---
t---
0
20 0
20
1.0
0
2. 0
"
"LL:=
I4
j
3.5
3.0
/15V'
I. 5
/
Initial Mach numbe r, M1
FIG. 3- 10. Shock angle ve rsus inlet Mach number and tu rning angle [1]. Curve~ above dashed li ne hold for M , < I and curves below ho ld for M z > 1. (After ShapIro[!].)
... ~
1.0
0.9
~:-("~
c 0.8
~
-"0 8
~ 0.7
~~
~
------ Gr
t------
~ ~" ~ ""::::: I ~L\, , I'---- 30~
,!2
ee 0.6 o
~
~ O. 5
Normal 4
s~:ck , ~ ~> ~
3
~-...-:: "-35"--
~
O. 2 y~
o.1
1.0
1.5
2.0 2. 5 Initial Mach' nurnber, M\
3.0
I I
V
1.,4
3.5
FIG. 3- 11. Stagnation pressure ratio ver sus inlet Mac h number , with turn ing angle as pa ramete r [I]. Curves above dashed line hold for M z > 1 a nd cu rves below hold for M , < 1. (After Shapiro [11.)
shoc~
1.0
1.5
/rV
5mlx or
-~--,
o.5
/
7V 1 7 ~ ~~~H'\~,:= __
Normal
-tw--
I-- f-..-
V
// V/ 2~ / / V/ f/ / ;;;/ / V
~\ t~1\
\
50 10 0
~~,
I.
V
/
V 1/
V
1/
11VI 0/1
Mach wa ve
Deflection angle 0= 00
5~~V V
I/ V lr
u
~
2.5
2.0
1.5
Dd1ection angle
0
.c
,~
10
1/ / 1/ VV VV
2. 5
E
----r:::: ~ ~
30
~,
61
I I
:
2.0 2.5 Initial Mnch !lumber, M\
: 3.0
3.5 )' = 1.4
FIG. 3- 12 . Exit Mach number versus inlet Mach number, with turning ang le as parame ter [1]. Curves above dash ed line cor respond to small (J and curves below corn. spond to large~. (After Shapiro [I].)
be determ ined from M
In
T 2 / T, and the condi tion
and M 1 and the d ownstream Mach num ber from M 211, lilt
= "2 /-
Alternatively, the above four eq ua ti ons, alo ng with the equa tion of sta te, can be reduced to a set of four independ e nt equati o ns relati ng the va riables M I , M 2, ",II'" p ,;p 1, fT , and 0 [I]. If any two variables are give n, for exam ple M 1 and the turni ng angle 0, then all others are de term ined and the downstream (or fina l) conditions can be founel in terms of th e upstream conditi o ns. The solution of these equations from Shapi ro [IJ is given in Figs. 3- 10, 3- 11 ancl 3- 12 fo r th e case f = l .4 and fo r various v:.1 1ucs of o. Axisym metric oblique shocks arc di scussed by Shapiro [3], who presents a comp lete solution for the case of th e conical shoc k as we ll as methods of treatment of the general axisymmet ri c shock problem .
62
STEADY ONE-DIMENSIONAL FLOW OF A PERFECT GAS
3- 6
References 1. SHAPI RO, ASCHER H., Tire Dynamics and Thermodynamics 0/ Compressible Flllid Flow Volume I. New York : The Rona ld Press Company, 1953; Chapters 1-8, 16 2. 'KEENAN , J. H., and J. KAYE, Gas Tables. New York: John Wiley & Sons, 1957 3. SHA PIRO, ASCHER H., op. cit., Volume H, Chapter 17
3-6
SHOCKS
63
5. A Mach 2 flow passes th rough an oblique shock as show n, a nd deflects 10°. A second obliq ue shock reflects from the solid wall. Wha t is the pressure ratio across the twoshock system? It may be assu med that there is no boundary layer nea r the wall; i.e., the flow is uniform in each of the regions bou nded by the shocks.
M =2
10° • ~
Refl ec ted
oblique shock ~ PROBLEM
Problems 1. Show that for a pure substance the stagnatio n pressure cannot increase for an adiabatic zero-wo rk process. During what physica l processes might it increase for a zero-wor k process? 2. Show that, for a perfect gas, the Mach number at the nose of the Fanno and Rayleigh lines is indeed unity. The nose is defined by ds/(rr = O. 3. A perfect fluid expa nds in a frictionless nozzle fr om stagnation conditions Po = 600 psia, T o = SOOooR , to ambien t pressure of 15 psia . If the expansion is isentropic, determ ine the followin g conditions a t the fin al pressure : (l) velocity, (2) Mach number, (3) temperatu re, a nd (4) area pe r un it ma ss flo w. How does the fi nal flow a rea compare with the th roat area for a give n mass flo w? The specific hea t ra tio 'Y is 1.4 and the molecul ar wei ght M is 30.
MI~ ~ f-:o,.--:=----- ----,=_ T
P:~
L----
2
M, T, PI
PROBLEM 4
4. Air fl ows through a cy lindrica l combustion c hamber of diameter 1 ft a nd lengt h 10 ft. StationsQ) andQ) are the inlet and ou tlet, res~cti velY .. The inlet stagna ti on temperatu re is SOooR and thel in let stagn at ion pressure IS 200 P SIa. (a) H the skin friction coefficient is C[ = 0.0040 (a pproximately uniform) and no comb ustion ta kes pla ce, wha t will the inlet Mach number be if it is subsonic and the exit Mac h number is un ity? What are t,he values of stat ic and stagnat ion pressures a t the exit? (b) If combustion takes place and the heat of reaction is 300 Btu per pound of mixture, neglecting fri ctional effects, and if the inlet Mach numbe r is 0.25, what is the exit Mach number? (c) If, in a simi lar problem, the effects of combust ion an d skin fri ction on ~he flow are o f the sa me order of magn itude, how might thei r simu lta neo us effects be estimated ? The flui d may be cons idered a perfect gas with 'Y = 1.4 a nd M = 29.
5
6. Air ente rs a constant-a rea duct at Mach 3 a nd stagnation cond itions 13000 R a nd 185 psia. In the duct it undergoes a frictionless energy-tra nsfer process such that the exit Mach number is unity. Consider two cases : (a) normal shock at in let to the duct, a nd (b) shockfree supersonic heating. Determine the stagnat ion temperat ure and press ure at the exit in each case. Is there any reaso n why the total energy transfer should d iffer (or be equal) in the two cases? 7. A uniform mixture of very sma l1 solid particles and a perfect gas expands adiabatica lly fro m a give n cha mber tempera ture through a given pressure ratio in a nozzle. The particles a re so small th at they may be considered to trave l with the loca l gas speed. Since the density of the solid material is much higher tha n the gas phase, the total volume of the pa rticles is negligible . The rat io of solid to gas flow ra tes is Ji., where o < J1. < 1. Consider two cases : (a) no heat tra nsfer between solid a nd gaseous phases, and (b) the solid particles have the loca l gas temperature at all points in the flow. Qualitatively, how does the presence of the so lid pa rticles affect the ve loc ity at the end of the expansion? Show that fo r both cases the mi xture is equivalent to the adia ba tic expansion of a homogeneous gas of different molec ular weight M and speci fic heat rat io 'Y . 8. A 30-ft3 rocket- propellan t chamber is filled with combustion gases at 60000 R and 1000 psia at the instant combustion ceases. If the throat area of the nozzle is 1 ft 2 esti mate the time it takes for the chamber pressure to drop to 100 ps ia. Assume tha~ at the insta nt combust i ~ ceases the chamber propella nt supply stops completely, a nd also that l' ~ 1.2 and M ~ 20. 9.
(a) Compare the work of compression pcr unit mass of ai r for both reversible ~~i~batic a nd reversible isothermal compress ion through a press ure ratio of 10 with InIttal conditio ns SOQoR and 14.7 psia. Is the importa nce of specific hea t variation large? Why should cool ing d ur ing compression red uce the wo rk ? . (b) In a n effort to provide a continuous cooling process during compression wate r IS sprayed int? a n airstrea m ente ri ng the axia l compressor of a gas turb ine ~ngine. The compressIOn may be taken to be reversible an d the evaporatio n rate is ap proximately.dw/ dT = k, where HI is the wa ter-a ir ma ss ratio , T is the mixture temperatu re, and k IS a constant. If w« I , the process may be simplified by consider ing the sale effect of the evapo ration to be e nergy trans ferred from the gas phase . Show how the fi na l temperat ure and the work of compression depend o n k.
4-1
THE VISCOUS BOUNDARY LAYER 2
4
"\
supercritica;/
1,,\~,SUb:itical
Boundary Layer
-2
Mechanics and
-3
Heal Transfer
4-1
65
THE VISCOUS BOUNDARY LAYER
Historically the development of boundary layer theory fo~lowed the development of mathematical solutions for flows assumed to be nonvlscou~. In th~ latter half of the nineteenth century-after Euler had formulated the ba~lc equatlOns of motion-Helmholtz, Kelvin, Lamb, Rayleigh, and other~ ap~hed them t.o a number of physical situations. By neglecting the effects of VISCOSIty they obtaIned elegant mathematical descriptions of various flow field~. .I~ was argued that a fluid like air, for example, which is so rarefied as to be InvlSlble, must ~ave negligible viscosity. However, with some exceptions, the so-called perfect-flUId theory disagreed strongly with the results of experiment.
(a) Perfect fluid
(b) Actual fluid, very low velocity
(c) Actual fluid, low velocity
FIG. 4-1. Flow patterns about a cylinder. One particular example, which came to. be kn~wn as d'Alembert's .paradox, attracted a great deal of attention. As an illustratIOn.' th~ flow of a flUId over a cylinder as predicted by nonviscous theory is shown In Fig. 4-:-1 (a). The theory predicts that the velocity and pressure fields are both s~mmetflcal about a plane norma I t 0 th e ups t re am velo city vectors ' Thus . the flUid , can . .exert no net forces (lift or drag) on the cylinder. Moreover , .th~s c?ncluslOn IS Independent o.f the shape of the body; if the fluid is truly invlscld, It can be shown mathematically 64
\ o
)(
I
I /
r---
r! - --
~.../ '/ 2
- - - Perfect-fluidtheo ry prediction -
-
- Measured results
FIG. 4-2. Pressure distribution on a cylinder. (After Flachsbart [11].) that for any body shape the net integrated pressure force upon the body will be zero. Of course, this conclusion directly contradicts experience. Everyone knows that it takes a finite force to hold a body immersed in a moving stream or to propel it through a stationary mass of fluid; hence the paradox. Observations of the actual streamline pattern around the cylinder reveal something like Fig. 4-1 (b), which suggests that the streamline symmetry is grossly disturbed. On the rear side of the cylinder the fluid seems almost stagnant when the stream velocity is quite low. For higher speeds, fairly violent oscillations take place, which affect the entire streamline pattern near the cylinder. Under certain conditions it is possible to observe a periodic shedding of vortices alternately from the top and bottom sides, as suggested by Fig. 4-1 (c). This downstream flow pattern has been called a vortex street. The drag of the actual fluid on the cylinder is largely due to the asymmetry of the pressure distribution on its surface. Figure 4-2 shows measured pressure distributions on a cylinder and , for comparison, the prediction of perfect-fluid theory. The symbol p stands for fluid density and the pressure and velocity far upstream are denoted by poo and uoo . It can be seen that two experimental pressure distributions may be measured (the difference between the two will be discussed later) and that both differ greatly from the theoretical result. The fact that the average pressure on the cylinder is much higher on the upstream hemi-surface explains most of the drag. In addition to this, the viscous force of the fluid on the cylinder imposes a drag force which, in this particular case, is considerably smaller than the pressure drag. For " streamlined" bodies, which have much less pressure drag, the viscous force is relatively more important. For quite a period the discrepancy between existing theory and physical results was unexplained, and a marked rift developed between the theoreticians, who continued to expound the perfect-fluid theory, and the engineering-oriented workers like Bernoulli and Hagen , who actively developed the empirical body of knowledge generally called hydraulics . The equations of motion which do take viscosity into account-the Navier-Stokes equations- were formulated but were
66
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
4-1
so very complicated that they could not, in general , be solved. It appeared exceedingly difficult to improve mathematically on the perfect-fluid theory except in a few very special cases. Then Prandtl (in 1904) made a major contribution to the solution of the problem by introducing the concept of the viscous boundary layer. . He showed that even for fluids of vanishingly small viscosity there is a thin regIOn near the wall in which viscous effects cannot be neglected. Since the velocity of the fluid particles on the wall is zero, there may be, under certain conditions, a region of large velocity gradient next to it. Then, even though the viscositY.M may be s.malI,. the shear stress on a fluid layer, T = McaU/ ay), may be large In that regIOn since au/ ay is large. This region is called the viscous boundary layer. ~ bound.ary layer of thickness 0 is indicated in Fig. 4-3. Outside t~e laye.r, the velocity gradient au/ay will be so small that viscous shear may be qUIte unlmportant, so that the fluid behaves as though it actually had zero viscosity.
FIG. 4-3. Boundary layer velocity distribution. Thus Prandtl showed that the flow field may be broken into two parts; a thin viscous zone near the waIl, and an outer zone where the nonviscous theory is adequate. This suggested the major simplification o~ the ~avier-Stokes e~uati~ns which is the basis of boundary layer theory. For fairly hIgh-speed flow, In WhICh the boundary layer is so thin that the pressure gradient normal to the wall is negligible, Prandtl and others showed how to treat the boundary lay~r mathematically for special cases, many of which have been solved in the past SIX~y y~ars. Given that a boundary layer exists on a body, it is easy to see qualitatively why the case of vanishing viscosity is fundamentally different from the case of zero viscosity . Figure 4- 2 shows that for zero viscosity a particle on ~h~ su:face streamline of the cylinder rises in pressure to the stagnation value when It Impinges upon the cylinder (8 = 0). Then the pressure falls to a minimu~ (~nd the velocity rises to a maximum) at the position 8 = 7r/2. From that pOint It travel~ to the rearward stagnation point (e = 7r), having just enough momentum to chmb the "pressure hill" and arrive at the stagnation point. . However small the viscosity, the viscous case is fundamentally different. The fluid near the wall has been slowed by viscous action. When a particle in the boundary layer reaches e = 7r/ 2 it will have reduced momentu~ so th~t only a small pressure rise will stop its forward motion, or actually send It movmg back-
THE VISCOUS BOUNDARY LAYER
4-1
--
/ / /
67
Compressor cascade
FIG. 4-4.
Subsonic diffuser
Rocket nozzle
Some practical devices in which separation may occur.
ward. For this reason the particle can no longer follow the contour of the wall. The accumulation of stagnant fluid on the back part of the cylinder deflects the outer streamlines, as suggested in Fig. 4-1 Cb). This in turn greatly modifies the wall pressure distribution, as shown in Fig. 4-2. When the streamlines in the vicinity of the wall cease to follow it, they are said to separate. This accumulation of stagnant or near-stagnant fluid and the resultant gross distortion of the streamlines means that the boundary layer behavior can have an important influence on overall flow behavior even though the quantity of fluid directly affected by viscosity is a small fraction of the total flow. Three separated flows of practical significance are shown in Fig. 4-4. Both the compressor cascade and the subsonic diffuser are designed so that the mainstream velocity decreases in the flow direction; hence the flow is against an adverse pressure gradient. If the pressure rise is too great the boundary layer will separate, creating regions of near-stagnant fluid as shown. Within rocket nozzles the pressure gradient is usually favorable, so that separation need not be a problem. However, under certain off-design conditions shocks can occur within rocket nozzles and the sudden pressure rise across the shock can cause separation. Separation and its consequences in these devices will be discussed in the following chapters where appropriate. Our purpose in this chapter is to establish the nature and cause of separation and to indicate how one might predict its occurrence (although this is seldom possible with purely theoretical methods). Separation is, of course, only one phase of boundary layer behavior. We shall be interested also in the behavior of non-separated boundary layers, since they are of importance to heat transfer and other phenomena. The classic picture of separation is illustrated in Fig. 4-5, which shows a series of velocity profiles in a boundary layer as it approaches and passes the separation point. As the pressure rises the free-stream velocity U falls, as the Bernoulli equation would predict. However, the effect of a given change in pressure is greater on the slow-moving fluid near the wall than on the free-stream fluid.
68
4-1
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
4- 1
THE VISCOUS BOUNDARY LAYER
69
cussed SUbsequently.) With these assumptions , the momentum equation requires that
1)1-______ u y
j
J
./
7
!
J
~
/
~/ ~/~~~ ~/
]
/ Streamlin e -Z-Which Ie ft ~___ the wall Recirculating flow ~~
-7
f--- x
=
-u du (boundary layer),
where, in keeping with boundary layer assumptions, the free stream and the boundary layer experience the same pressure change. Equating the two expressions for dp and defining the ratio a = u/ U, there results
J
/
p
p
---1
J
dp
dp = _ U dU (free stream),
da = _
d~ (~ -
pU2
(
2 )
.
a
Hence if the pressure rises (dp > 0), a decreases and the boun.dary layer fluid is decelerated more than the free stream. Noting that -dp/ pU 2 is simply dU/ U, this expression can be integrated (for constant density), yielding
Separation point
FIG. 4-5.
A simplified picture of the development of separation.
I -
The fluid near the wall can be slowed to a stagnant state and then be made to flow backward if the adverse pressure gradient continues. The separation point is generally defined as that point where a streamline very near the wall leaves the wall. In two-dimensional flow this corresponds to the point at which the velocity profile has zero slope at the wall. That is ,
(au) ay
=
y=O
°
ai
where the subscript 1 refers to an initial condition. The thickness 0 of the boundary layer can be found from the equation for continuity which requires that
or, again for constant density, OIl_Ul •
(two-dimensional separation).
OIU
Figure 4-5 suggests that, past the separation point, there is a region of recirculating flow. In practice this region is highly unstable. It may be quite difficult to determine analytically, since once separation occurs there is a strong interaction between the body of stagnant fluid and the free stream. Often there results an adjustment of the free-stream geometry such that the adverse pressure gradient is reduced to practically zero, in which case there is no force to drive the recirculating fluid. The effects of pressure gradient on the shape of the boundary layer, which are shown in Fig. 4-5, U may be explained in a simplified manner as follows . Let the boundary layer be replaced by a hypothetical discontinuous flow shown in Fig. 4- 6, in u which the continuous variation in velocity normal J to the wall is replaced by two step changes. Fur~ ther in order to focus on the effect of pressure gradient, let us for the moment neglect shear FIG. 4-6. Discontinuous stresses. (The effects of shear stresses will be dis- boundary layer. ~
~
~
~
T
The two quantities a and 0 are plotted against U/ U l in Fig. 4-7 for the particular case al = l It may be seen that this boundary layer separates (a -70, 0 -7 U) ) when the free-stream velocity has decreased by less than 14%. With this picture of the relatively large effect of a given pressure rise on the more slowly moving fluid, it would appear that any pressure rise applied to a real boundary layer would cause separation since, if the velocity goes continuously to zero at the wall, there will always be some fluid with insufficient momentum to negotiate the pressure rise. However, this tendency toward separation is resisted ~p to a point, by visc?us stress. That is, ~he slowly moving fluid can be "dragged'; hrough a pressure flSe by the faster flUid farther from the wall. An important ~etho~ of reducing. the separation tendency is to enhance this "drag" force by InCreasmg the effective shear stress through turbulent motion . The mention of turbulent motion introduces another important aspect of boundary layer behavior. In 1883 Reynolds showed that under certain conditions dye particles injected into a pipe flow moved smoothly along streamlines without late~al fluctuations . In the same pipe at higher velocities, this smooth laminar motLOn broke down into turbulent motion consisting of rapid random fluctua-
a 4-1
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
70
1.0,----- --
t -/
5 4
\ \
\_1
I
o0.7
( 0.8
0.3
....
0.2
-
=0
pDD. j..L
0.1
1.0
V
00
u
V
VI
FIG. 4- 7. "Growth" of step boundary layer near separation for a 1 = t·
FIG. 4-8. Comparison of laminar and turbulent velocity profiles.
tions superposed on the mean motion. After the flow had become t~rbulent the dye lines rapidly mixed with the surrounding fluid, the macroscoplC tu.rbul~nt mixing process being at least an order of magnitude faster than molecular diffusIOn (the only agency for mixing in laminar flow). . . The same transition behavior takes place in the boundary layer, resulting In a significant reduction in its tendency to sep~rate. The reaso~ for .this is that the turbulent fluctuations constitute a mechanism for transportmg high momentum from the outer part of the layer to the region near the wall.. The effect is ~s if there were an increased shear stress or, as will be shown, an Increased coefficlent of viscosity. This tends to raise the wall shear stress, but it also means that the turbulent boundary fluid can climb higher up the " pressure hill" than the fluid in the laminar layer. A comparison of typical turbulent and laminar boundary layer velocity profiles is indicated in Fig. 4-8. The fact that the turbulent layer tends to have much higher momentum (and shear stress) near the wall gives it an increased resistance to sep.arati.o~. As a method of specifying the resistance of a boundary layer to separatIOn, 1t 1S useful to define a pressure coefficient C p as f::.p
Cp
71
geometry is completely specified (far from the entrance) by pipe diameter D and wall roughness. The pertinent fluid properties are density p , viscosity j..L,and the average velocity D. These variables can be arranged in a dimensionless form known as the Reynolds number, Re, so that Re
0 0.9
THE BOUNDARY LAYER EQUATIONS
y Ii
ex
ii
',ii i
--------,
0.4
fa -
-
0.5
4-2
=
lxPU2'
where U is the free-stream velocity at the point where the pressure begins to rise . and f::.p is the total increase in pressure up to the point in question. For an approximate comparison of cases in which the adverse pressu~e gradIent is not abrupt (i.e., where the shear stress has time to act) it may be saId that the laminar boundary layer will support a pressure rise given by 0.15 < C p < 0.2, whereas the turbulent boundary layer can tolerate without separation a pressure rise corresponding to 0.4 < C p < 0.8, depending on the Reynolds number and the upstream pressure distribution of the flow. . .' The onset of turbulence, or transition as it is called, IS dependent In a given device on fluid properties and velocity, the general or upstream turbulence level, and the geometry of the particular device. In the case of pipe flow, for example,
Reynolds found that for "smooth" pipes (commercial pipes and tubing are usually "smooth") and very quiet inlet conditions, the transition process could in fact be correlated with a critical Reynolds number below which the flow was laminar and above which it was turbulent. More generally it is found that transition in any series of geometrically similar devices can be correlated with the Reynolds number, where D is replaced by some characteristic length Land D by some characteristic velocity U. Thus Re
=
pUL. j..L
Obviously, a critical value for Re can have meaning only within the geometry for which it was defined . With these facts in mind, the explanation of the two pressure distributions observed on the cylinder of Fig. 4-2 becomes straightforward. For reasonably low velocities (below a "critical" Reynolds number) the boundary layer remains laminar and separates readily on the back of the cylinder' before much pressure rise has taken place. At the critical Reynolds number, the boundary layer becomes turbulent and is able to flow considerably further around the back of the cylinder before separation . As it does so it rises in pressure, thus reducing the discrepancy between inviscid and actual pressure distributions. Since the drag on a blunt body such as a cylinder is primarily a pressure force (as opposed to skin friction), the onset of turbulent flow in this case is actually accompanied by decreased drag. In most practical fluid machines to be discussed in this book the bound ary layers a.re turbulent, and so long as this is true the influence of Reynolds-number variations on overall performance is usually minor. However, under extreme conditions, as for exampl e at very high altitudes, the Reynolds number can become low enough that portions of the flow revert to laminar flow. There usually follows a rather substantial reduction in efficiency of performance, due to the increased tendency toward separation.
4-2
THE BOUNDARY LAYER EQUATIONS
T~e general equations for the flow of Newtonian viscous fluids are called the
Nav~er-Stokes e~uations [1,2].
Because of their mathematical complexity, it is pOSSIble to obtam exact solutions of them for only a few physical situations. I many cases, .howeve:, it is legitimate to make simplifying approximations an~ thereby obtaIn solutions whIch have a useful range of validity. The boundary
72
4-2
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
4-2
THE BOUNDARY LAYER EQUATIONS
73
y
Table 4-1
u
_ _ _+-__TX_+ ..
aTx
Surface
ay dy
Mass flux
(}),---- -----,Cll I I
I I I ~ I
I~ I
I I p
pV
I I
4-9. Small control volume for development of boundary layer equations. FIG.
~ x
+-__
_ _+-1
Control volume
+ aTy d
p( u + ~ dX)
+ fJ! dx ax
-----J(D
layer equations are much more tractable than the Navier-Stokes equations. They may be derived by neglecting certain terms in the Navier-Stokes equations which can be shown to be quite unimportant in boundary layer flows. The method is developed by Schlichting [I J in careful detail. In the following discussion , equations will be developed less rigorously, but more directly, for the special case of the steady, two-dimensional, incompressible boundary layer. Shapiro [3J uses this approach to develop equations for compressible flow, and includes a good discussion of the range of validity of the simplifications. The object in view is to lay a foundation for a quantitative description of both laminar and turbulent boundary layers, with particular emphasis on skin friction, separation, and heat transfer. Although the solutions of the boundary layer equations which are available are seldom directly applicable to the complex flows in real machines, they at least illustrate the forces at work and the behavior typical of the boundary layer. Consider a very small control volume within a boundary layer, as shown in Fig. 4-9. Velocities in the x- and y-directions are u and v, respectively, while the shear stresses in these directions are Tx and Ty • To apply the continuity and momentum equations to this control volume, we shall need to know the mass and momentum flux across each surface and the force (shear and pressure) acting on each surface. The outward fluxes of mass and x-momentum are shown in Table 4-1. Since the flow is taken to be two-dimensional, no fluid crosses the surfaces parallel to the xy-plane. The continuity equation for steady, two-dimensional, incompressible flow requires that p (u
or
+ :~ dX) dy
-
+ p (v+ ~~ dY) ~':!. + av = o.
pu dy ax
ay
dx -
+ ~: dX) dy
CW)
p (u
@@
-pudy
COO)
p (v
@)(D
-pvdx
+ :; dY)
(u + ~: dX)
2
dy
dx
p (v
+ :;dY)
(u
+ ~;dY) dx
-pvu dx
The momentum equation in the x-direction, using Table 4-1 to obtain the net outward momentum flux, is (applying Eq. (2-4»
L Fx =
p (u
+ ~~ dX) 2 dy -
2
pu dy
+ p (v + :; dY )
(u
+ ~; dY ) dx -
puv dx.
Expanding and disregarding second-order terms, we have
L Fx = p (2U axau + u ayav + v ayau) dx dy. Using the continuity equation, we may simplify this to
L Fx
= p (u au
ax
+ v au) dx dy. ay
~ince only pressure and shear forces act on the control surface, the force summatiOn may be written, in accordance with Fig. 4-9, as
p pdy - (p + aax dX) dy + (Tx + aTX) ay dy dx (-
Tx dx
~~ + ~?) dxdy.
Thus the momentum equation may be written
X
(4-1)
p
_pu 2 dy
p (u aau
pu dx = 0
Flux of x-momentum
+
~p. + aTayx .
u v aa ) = _ y ax
Th~ shear ~tress T developed in a Newtonian fluid is proportional to the rate of sheanng straIn. For the particular case of the boundary layer, in which there is a
74
4-2
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
large velocity gradient close to the wall, the shear stress is very nearly equal to
4-2
THE BOUNDARY LAYER EQUATIONS
Thus this equation may be written
f
au ay
f..t-'
h
o
where f..t is the coefficient of viscosity. Using this definition, the momentum equation for the x-direction can be written
75
(u
au ax
h
au + v oy
_
dV) U dx dy
f
=
0 /I
2 a u ay2 dy.
Integrating the right-hand term and noting that the wall shear stress is given by TO
=
f..t
(au) , ay y=o
while au/ oy is zero outside the boundary layer, we can write (for constant As a consequence of the thinness of the boundary layer, pressure changes normal to the wall may be neglected inside the layer, so that it may be assumed that p is a function of x only. Then the momentum equation may be written
h
u ou / o ( ax
+ v au
v dV) d
_
oy
dx
= _
y
/I)
TO.
P
I ntegrating the second term in the integral by parts, we have (4--2)
where /I =::0 f..t/ p is the kinematic viscosity. The momentum equation as derived here assumes a fluid of constant viscosity and constant density. Although these conditions may be satisfied within many low-speed laminar boundary layers, we shall see that the "effective" viscosity in turbulent flow may vary strongly through the boundary layer. Further, p and f..t could vary as the result of temperature variation within the boundary layer, due either to high heat-transfer rates or, in compressible flow, to the variation of static temperature in response to velocity variation. In such cases a complete set of equations must include at least an energy equation and an equation of state. However, even in cases not requiring the simultaneous solution of an energy equation, Eqs. (4-1) and (4-2) can present considerable difficulties in the solution for u and v (where p is assumed a known function of x). It is desirable, indeed necessary in most cases, to utilize approximate methods for solving the boundary layer equations. An extremely useful approximation, the momentum integral method, has been developed and widely applied. The object of the method is to find a solution which will satisfy the integral of the momentutn equation across the boundary layer even though it may fail to satisfy the equation at particular points. The integration of the momentum equation (4-2) may proceed as follows:
foh(u auax + v ayau) dy = fh 0
1 dp
-
P dx dy +
fh 0
lJ
a2u ay2 dy,
where h is an undefined distance from the wall outside the boundary layer. the free stream, I dp UdU. dx dx
P
In
h hVaudY f o oy
vul
=
j'h u~1!.dY.
_ 0
From the continuity equation, v at y
= h is given by
h
v
/
ay
0
av -dy o ay
h
au
= - /0 -dy. ax
Thus
{ h au v ay dy = -
} 0
v
fl< au 0 ax dy
+
fh 0
au u ax dy.
The momentum equation may then be written
l< fo
(2U ax au
-
v au
ax
-
v dV) dx
dy
=
TO
p
or, rearranging,
fo h
a~ [u(V -
u)] dy
+ ~~ fo h (U
- u) dy =
+ ~o.
(4--3)
Both the integrands are zero outside the boundary layer, so h can be indefinitely large. . Equation (4-3) is arranged in this particular form because of the physical sig?Ificance (and the common usage) of the individual integrals. In the second Integral, (U - u) is the mass flux defect (actually I/p times the defect) that occurs as the result of the deceleration of the boundary layer fluid. In the case of flow within a duct, for example, if there is a mass flux defect near the walls there must be an increased mass flux within the free stream. This effect on the free stream would be the same if there were no boundary layer but the duct walls were dis-
76
BOUNDARY LAYER MECHANICS AND HEAT TRANSFER
4-3
4-3
lAMINAR BOUNDARY lAYER SOLUTIONS I. 0
placed inward an amount 0* so that Vo*
=
10'"
(V -
0.8
u
U 0.4
0*
=
faoo
(I -
~)
e = faoo ~ (I - i ) dy.
(4-4a) (4-4b)
I
R
1.0
=
Ux v
o 1.08 X 10 5 o 3.64 XIO·; /:,. 7.28x I05
2.0
3.0
4.0
5.0
5.646.0
7.0
q=y!!Z.
~~ Velocity distribution in the laminar boundary layer on a flat plate at zero InCidence, as measured by Nikuradse [IJ . ~IG. 4-10.
The simplest flow is that over a flat plate where dpl dx is zero. In this case, Eqs. (4-1) and (4-2) reduce to
With these definitions, the momentum integral equation, Eq. (4-3), may be written
u au
ax
+ v au ay
=
with the boundary conditions
or
(4-5) As it stands this equation applies to both laminar and turbulent boundary layers. The solution, say for a given Vex), requires knowledge of the relationship between 0* and e, which may be found if an approximate velocity profile shape is assumed (0* 1 e is often called the shape factor H), and the skin friction TO is known. Note that shear stress is evaluated only at the wall where, even in turbulent flow, it is given by M(aul ay), since turbulent fluctuations must go to zero at the wall. 4-3
i
I
o o
.
dy,
/
V
( 00
ve=}o u(V-u)dy. Here e is called the momentum thickness of the bounda ry layer. The two terms are then defined as follows (for incompressibldow):
/1'"
0.6
0.2
~
/ 11
u) dy.
With respect to mass flux then, the boundary layer has the effect of making the thickness 0* unavailable for free-stream flow ; 0* is called the displacement thickness of the boundary layer. The first integral can be treated in a similar ma nner. Note first that since the limits of integration are not functions of x, the partial differentiation al ax can be carried outside the integral where it becomes d(dx. In this case we interpret u as l i p times the actual mass flu x, while (V - II) is the momentum defect per unit mass of the boundary layer fluid . Hence the integral is the momentum defect in the boundary layer which, as above, could be accounted for by the loss of free-stream thickness e such that 2
77
y = 0:
u
y= 00:
u = V = constant.
v
=
=
0,
A stream function t/; may be defined such that v
_ at/;.
=
ax
The convenience of this form is that it automatically satisfies the continuity equation so that the above two equations reduce to
LAMINAR BOUNDARY LAYER SOLUT:ONS
A variety of exactt solutions of Eqs. (4-1) and (4-2) are available, two of which will be mentioned here .
(4-6) This equation may be considerably simplified by substituting
t Exact in the sense that they do not neglect anI of the terms in the boundary layer equations. Most solutions are, however, expressed in terms of series expansions.
1/
= y ".,! V /vx
and
if;
= vvxU /(1/) ,