You are on page 1of 310
THERMODYNAMICS AN ADVANCED COURSE WITH PROBLEMS AND SOLUTIONS RYOGO KUBO University of Tokyo in cooperation with HIROSHI ICHIMURA TSUNEMARU USUI NATSUKI HASHITSUME Tokyo Institute of Technology Nagoya University Ochanomizu University 1968 NORTH-HOLLAND PUBLISHING COMPANY - AMSTERDAM JOHN WILEY & SONS, INC. - NEW YORK (Interscience Publishers Division) © NORTH-HOLLAND PUBLISHING COMPANY - AMSTERDAM 1968 No part of this book may be reproduced in any form by print, photoprint, microfilm or any other means without written permission from the publisher. Library of Congress Catalog Card Number 67-26462 THIS BOOK WAS ORIGINALLY PUBLISHED IN JAPANESE BY ‘THE SHOKABO AND COMPANY, TOKYO PUBLISHERS: NORTH-HOLLAND PUBLISHING COMPANY - AMSTERDAM. SOLE DISTRIBUTORS FOR U.S.A. AND CANADA? INTERSCIENCE PUBLISHERS, A DIVISION OF JOHN WILEY & SONS, INC, - NEW YORK PRINTED IN THE NETHERLANDS PREFACE TO ENGLISH EDITION The original text of this volume is that of the first four chapters of the book “Problems and Solutions in Thermodynamics and Statistical Mechanics” itself one of the “University Series” published by the Shokabo Publishing Company. Although the last part of the original book was translated first and published as “Statistical Mechanics-An Advanced Course with Problems and Solutions” (North-Holland Publ. Co., 1965), the present vol- ume should be regarded as the first volume of the whole work. The English title is somewhat different from the original one and it seems to have caused some misunderstanding of the authors’ intention of writing this book. This is not the usual type of textbook, but it is aimed primarily to help the student to study the subject by thinking for himself and by work- ing problems by himself. The solutions of the problems, of course, can be used in various ways. The lectures, entitled the fundamental topics, are therefore meant to be as concise as possible to make it convenient for the reader to refer to them repeatedly. Much of the material usually treated in textbooks is included sometimes in disguised form, in the examples or prob- lems. In this respect this book is selfcontained. However, it may be recom- mended students to use a standard textbook together with the present book. Most of the translation was done by Dr. Hajime Seki, IBM Laboratory in Yorktown Heights, N. Y. and some parts were done by the original authors. As the editor of the original Japanese edition and of the English edition, I wish to express my deep appreciation to Dr. Seki, and also to Professor Bernard Springer of the University of Southern California and Professor Donald C. Worth of International Christian University, Tokyo, who kindly took the trouble of reading through the manuscript and helping us with linguistic difficulties. My thanks is also due to Mrs. N. Hashitsume and Mrs. M. Suzuki who helped us in preparation of the manuscript. 1966 Ryoco Kuso PREFACE TO JAPANESE EDITION Thermodynamics and statistical mechanics are indispensable tools in studying the physics of the properties of matter. Statistical mechanics, together with quantum mechanics, provides a foundation for modern physics which aims at the thorough understanding of physical phenomena from the microscopic viewpoint of atomic physics. Fundamental knowledge and training in statis- tical mechanics are therefore of vital importance not only for students studying the physical properties of matter but also for those who study nuclear physics or even astrophysics. Outside the realm of physics, its im- portance is rapidly penetrating into chemistry, biology and into those vast areas of technology which owe their growth to the advances in modern physics. Thermodynamics belongs completely to classical physics and is some- times regarded as unimportant by students of physics who are over-occupied in learning modern physics. Even for students in chemistry, the present is different from the time some decades ago when physical chemistry was al- most nothing but chemical thermodynamics. However, it must be stressed here that the usefulness and unique significance of thermodynamics as a fundamental science remain as basic today as they were in the latter half of the last century. Thermodynamics teaches us the value of a phenomenological ‘approach. It avoids explicit use of physical images or models sich as atoms and molecules. Instead it deals with relations between somewhat abstract quantities such as energy, entropy, free energy and so forth. Admittedly it does not give intuitive pictures as atomic theories do, which is one of the reasons why students find it difficult to gain sufficient understanding and familiarity to use thermodynamics in real problems. But the simplicity of the logic of thermodynamics sometimes makes us see more clearly into the nature of the basic physics of a given problem from very general principles. This is the great advantage of a phenomenological approach. Obviously, however, it is impossible to explore more deeply the under- lying atomic processes in a given physical phenomenon if we confine our attention to thermodynamics. Such progress is made possible only by quantum mechanics and statistical mechanics. Statistical mechanics provides PREFACE TO JAPANESE EDITION vit us with a means to link the physical laws of the microscopic world to those of the macroscopic world. Without close cooperation with statistical mecha- nics, quantum mechanics itself would not be able to represent the physics of the real world. In these sense, statistical mechanics is indispensable as one of the keystones of modern physics. Like any other science, statistical mechanics cannot be mastered easily just by learning its principles once. One has to think by oneself a great deal before one grasps the way in which to use the statistical approach in one’s think- ing, and to apply statistical mechanics to real physical problems. In statis- tical mechanics and in thermodynamics, there are certain aspects which are quite different from other fields of physics. We often meet students who find difficulty in mastering thermodynamics or statistical mechanics, lacking confidence in applying it to real problems, although they know the prin- ciples. Such difficulties are due to insufficient and inadequate training. The purpose of the present book is to provide a guide for students studying and acquiring facility in thermodynamics and statistical mechanics. Thus it contains fundamental topics, examples and a fairly large number of problems with complete solutions. The fundamental topics are rather condensed, but still they cover all of the points which are basic. This book is meant to be read- able without reference to other textbooks. By reading through these topics only, one would be able to obtain fundamental knowledge of thermo- dynamics and statistical mechanics. The examples are partly to supplement the fundamental topics, but they are primarily meant to show the reader how the principles are applied to physical problems. The problems are classified into three groups, A, B and C, in order of in- creasing difficulty. If a reader has enough time he may go through all problems in each chapter. But, if not, it is recommended that he studies first the prob- lems in group A throughout the whole book and then later comes back to try Band C. By just finishing group A problems, he will find himself to have ob- tained a much better understanding of physics. The number of group A problems is fairly large, so that he may even select about half of these and come back later to the other half. The subjects in the fundamental topics and examples which are marked by * are not needed in solving problems in group A. In this bookf, thermodynamics and statistical mechanical problems are + The reader is reminded that this text is a translation of the Preface to the original Japanese edition, in which thermodynamics and statistical mechanics are contained in ‘one volume. vin PREFACE TO JAPANESE EDITION mostly limited to those of equilibrium states. It might be desirable to include kinetic methods and extensions of thermodynamics and statistical mechanics which apply to non-equilibrium problems. We had, however, to content our- selves in treating such topics in a limited way only in the last chapter (Chapter 6 of ”’Statistical Mechanics”). This is because the whole volume had become much larger than the original plan and also because such non- equilibrium problems are certainly somewhat advanced. As mentioned previously in this preface, quantum mechanics is the funda- mental dynamics of the microscopic world. In this sense, statistical mechanics ought to be essentially quantum-statistics. However, since the present book devotes itself to clear understanding of the nature of statistical considera- tions, only an elementary knowledge of quantum mechanics is required in studying problems in groups A and B. Therefore, even those students who are not specializing in physics but have only an elementary background in quantum mechanics will not find any serious difficulty in starting to study this book. What is most important in studying a physical problem is to grasp it as a problem in physics. Mathematical manipulations may sometimes be tedious and sometimes may require specialized techniques. Training in mathematical methods should not be ignored, but it would be a serious mistake if one was to be dazzled by the mathematics and to forget the physics. Teachers often meet students’ papers in which the student seems to be in no doubt about the numerical answers although they are in error by two or three orders of magnitudes or are dimensionally incorrect. Professor H. Nagaoka (a pioneer physicist in Japan) was carrying out calculations on a blackboard in his class. He changed the sign of his answer saying “It is plus rather than minus. Isn’t it?” Mathematical calculations may very often be in error. A physical mind is very important, for this can give you the right sign even when your calculation betrays you. An answer obtained by calculation is in many cases easily understood, at least qualitatively. It may not be guessed before making calculations, but one should not forget to think it over again in order to see if one can see some physical meaning contained in it. Such remarks are not given in each solution of the problems, so that we should like to empha- size here the importance of such reasoning. Here and there between the pages some comments ¢ are inserted under the title “Divertissement”’. While giving seminars to students we sometimes take a rest to drink a cup of tea and chat. We hope that the reader will spare a few + These are revised in this English edition, PREFACE TO JAPANESE EDITION 1x minutes at these spots to listen to a chat from the authors, drinking tea or coffee or just smoking. The fundamental topics were mostly written by R. K. Examples and prob- lems were selected after repeated discussion by all the authors. The final check of the solutions was made by R. K. and the whole design of the book was made by N. H. The authors would appreciate it if readers would kindly point out any mistakes which may have escaped our notice. Five years have passed since this book was originally planned, and two years since we started actually to write it. The undertaking proved to be much more difficult than we anticipated. The authors are particularly grate- ful to Mr. K. Endo, editor of Shokabo Publishing Company, for his continual encouragement and help. January, 1961 RyoGco Kuso CONTENTS CHAPTER 1. THERMODYNAMIC STATE AND THE FIRST LAW OF THERMODYNAMICS Fundamental Topics § 1.1 The Objects of Thermodynamics § 1.2 The Concept of Thermal Equilibrium (The Zeroth Law of Thermodynamics) § 1.3 Thermodynamic Contact § 1.4 Thermodynamic Quantities § 1.5 Processes of Change . § 1.6 The First Law of Thermodynamics § 1.7 Heat and Enthalpy . §1.8 The First Law of Thermodynamics Arplied to Infinitesimal Processes . $1.9 Temperature . § 1.10 Heat Capacity, Specific Heat § 1.11 Equation of State . § 1.12 Change of Independent Variables . Examples . Problems . Solutions . CHAPTER 2. THE SECOND LAW OF THERMODYNAMICS AND ENTROPY Fundamental Topics § 2.1 Reversible and Irreversible Process § 2.2 A Lemma (Carnot Cycle) § 2.3. The Second Law of ‘Thermodynamics § 2.4 The Efficiency of a General Carnot Cycle. §2.5 Absolute Temperature . §2.6 Clausius’ Inequality for an Arbitrary Cycle §2.7. Entropy §2.8 Additivity of Entropy §2.9 General Expression of the Second Law 61 62 63 63 65 69 70 4 14 CONTENTS § 2.10 The Direction of an Actual Change § 2.11 Maximum Work and Minimum Work Examples . aaa Problems . Solutions . CHAPTER 3, THERMODYNAMIC FUNCTIONS AND EQUILIBRIUM CONDITIONS Fundamental Topics § 3.1 Thermodynamic Functions . § 3.2 Legendre Transformation . § 3.3 Gibbs-Duhem Relation . § 3.4 Definitions of Thermodynamic Quantities and Thermodynamic Relations. . §3.5 The Third Law of Thermodynamics (Nernst- -Planck Theorem). § 3.6 Equilibrium between Two Systems Loe § 3.7 Equilibrium Condition for a Given Environment § 3.8 Inequalities of Thermodynamics §3.9 Le Chatelier-Brown’s Principle. Examples . Problems . Solutions . CHAPTER 4. PHASE EQUILIBRIUM AND CHEMICAL EQUILIBRIUM Fundamental Topics § 4.1 Phase . §4.2 Equilibrium between Different Phases of a Pure Substance . § 4.3. Surface Tension . § 4.4 Equilibrium of a Many- Component System it in Different Phases §4.5 Gibbs’ Phase Rule - 5 §4.6 The Chemical Potential of a Gas . : § 4.7 Chemical Potential and Saturation Vapor Pressure. of Liquids and Solids §4.8 Dilute Solutions. . . §4.9 Ideal Solution (Ideal Solid-: -Solution) and Regular Solution. § 4.10 Expressions for Concentration of Solutions . . § 4.11+Activity and Activity Coefficient § 4.12 Solution of Strong Electrolytes. xI 75 11 8 85 95 133 136 137 137 140 140 141 143 145 147 159 166 190 190 192 193 195 197 199 200 201 202 202 203 xm CONTENTS § 4.13 Chemical Equilibrium : § 4.14 Thermodynamics of Electrochemical Cells § 4.15 Application of the Third Law of Thermodynamics . Examples . Problems . Solutions . DIVERTISSEMENTS Chapter I 1. Founders of the first law of thermodynamics 2. Why do we have winter heating? . 3. Nicolas Leonard Sadi Carnot . Chapter 2 4. Absolute temperature . 5 5. International practical temperature scale. 6. Axiomatic foundation of thermodynamics 7. Kapitza’s helium liquefier . Chapter 3 8. On the names of thermodynamic functions . 9. Thermodynamic mnemonic diagrams Chapter 4 10. Equilibrium of heterogeneous substances . 11. Classification of phase transitions 12. Thermodynamics and life 205 206 209 210 228 11 36 65 71 128 147 161 194 196 249 CHAPTER | THERMODYNAMIC STATE AND THE FIRST LAW OF THERMODYNAMICS This chapter deals with the concept of a thermodynamic state and with those concepts related to the first law of thermodynamics. Students of thermodynamics must have a clear understanding of the physical meaning of such concepts. It should be noted, in particular, that work and heat are not functions of state but are quantities which are determined by specifying the process of change. Fundamental Topics § 1.1. THE OBJECTS OF THERMODYNAMICS Some of the basic concepts and terms used in thermodynamics are listed with explanations. In order to avoid unnecessary confusion, not all the definitions given here have been made rigorous. When necessary, more rigorous definitions are given later. System: Systems which are the objects of thermodynamics are macroscopic entities called systems with extensions in space and time which are accessible to normal processes of measurements. Such a system may consist of a great number of material particles (e.g. molecules, atoms, electrons, etc.) or of field quantities such as electromagnetic fields. In either case, they are dynam- ical systems containing an extremely large number of degrees of freedom. Systems composed of only a small number of degrees of freedom are not the object of thermodynamics. Environment: If part of a whole system is chosen as the object of our observation, the remaining part is its surroundings. The surroundings may be abstracted as an environment which defines certain conditions imposed on the system of interest (such as constant temperature, pressure, chemical potential, etc.). Isolated system: An independent system which has absolutely no inter- action with its surroundings is an isolated system. Closed system: A system which has no material exchange with its surround- ings is a closed system. Open system: A system which has interchange of material with its surroundings is an open system. 2 ‘THERMODYNAMIC STATE {Ch.1,§3 § 1.2, THE CONCEPT OF THERMAL EQUILIBRIUM (THE ZEROTH LAW OF THERMODYNAMICS) The thermal equilibrium state of an isolated system: Regardless of the complexity of its initial state, if an isolated system (e.g. a gas surrounded by a wall impervious to heat) is left standing, the system eventually comes to a final state which does not change. This final state is called the thermal equilibrium state. Microscopically the material particles still continue their complex motion, but macroscopically the thermal equilibrium state is a simple state defined by a few parameters such as temperature and pressure. Thermal equilibrium of two systems; When two isolated systems A and B are brought into contact with one another the total system A+B eventually comes to thermal equilibrium. It is then said that A and B are in thermal equilibrium with each other. The two systems, A and B are themselves also separately in thermal equilibrium. The equilibrium is not disturbed when the contact between the two is broken and is then restored after some time. Therefore two systems, A and B, even when they are not in contact with each other, may be said to be in thermal equilibrium with each other (A~B) if there is no change when they are brought into contact. The Zeroth law of thermodynamics (transitive law of thermal-equilibrium): If A and B are in thermal equilibrium in the above sense and B and C are also in thermal equilibrium, then A and C are in thermal equilibrium: A~B,B~C+A~C. (1) This empirical law is called the Zeroth law of thermodynamics. Thermodynamic state or state: In a narrow sense this term is used in the same sense as thermal equilibrium state, but in general a whole system is considered to be in a thermodynamic state if the various parts are in thermal equilibrium, i.e. in Jocal equilibrium, even when the system as a whole is not ina thermal equilibrium state. For example, if, in a system composed of two bodies A and B, each part is at a different temperature T, and Ty, it may be said that the entire system is in a state which is specified by (Ts, Tp). § 13. THERMODYNAMIC CONTACT Contacts which allow the following interactions between thermodynamic systems are called thermodynamic contacts. 1, Mechanical interaction: When one system does work on another system by means of mechanical or electromagnetic forces, the interaction is called mechanical. 2. Thermal interaction: An interaction which results in the exchange of Ch. 1, §4] ‘THERMODYNAMIC QUANTITIES 3 energy in the form of heat transfer by such means as thermal conduction and thermal radiation is called thermal interaction. A wall which cuts off thermal interaction is called an adiabatic wall. 3. Mass interaction: An interaction which involves the exchange of matter between two systems is called mass interaction. A semi-permeable membrane is an example. Idealizations of 2 and 3 are interactions weak enough not to alter the properties of the systems yet strong enough to result in effective contacts during the process of observation. If the environment is considered as a source which acts on the system under consideration in the way mentioned above, it is called respectively, a work source, a heat reservoir (heat bath) or a particle source (or reservoir). A source or reservoir is considered to be sufficiently larger than the system so that the source itself remains in a given equilibrium state regardless of the effect it exerts on the system through its contact with it. § 1.4. THERMODYNAMIC QUANTITIES State quantities: Physical quantities which have a definite value for each thermal equilibrium state of the system are thermodynamic quantities, or state quantities. Examples: temperature, pressure, internal energy, enthalpy and entropy. State variables or thermodynamic variables: In a narrow sense these are the thermodynamic quantities described above. In the broader sense of the thermodynamic state they denote the physical quantities which are defined by local equilibrium. If a set of independent variables is properly chosen so as to be necessary and sufficient to define a state, the other state quantities are their functions. The number of independent variables for a thermal equilibrium state of a system is determined empirically. Intensive quantities and extensive quantities: If a system in a thermal equilibrium state is divided into parts by inserting walls, each part stays in equilibrium. Therefore an equilibrium state of a uniform system is an internal property and is determined by state quantities which are independent of the extension or size of the system. Such quantities are called intensive quantities. Examples: temperature, pressure, chemical potential. On the other hand, state quantities which vary in proportion to the extension or mass of the system, when it is divided without changing the equilibrium state, are called extensive quantities. Examples: mass of components, energy, entropy. Internal variables and external variables: In some cases, thermodynamic variables are classified into the two categories: internal and external variables. 4 ‘THERMODYNAMIC STATE [Ch. 1,85 The external variables define the state of the surroundings. For example, the position of the piston in a cylinder containing a gas, and the strength of an electric field or a magnetic field acting on a system can be thought of as external variables, However, these may be considered as internal variables if the piston or the field sources are included in the system. Thus the dis- tinction depends on where we draw the line between our system and the surroundings in contact with it. This should be made clear in order to avoid confusion, particularly in the case of mechanical contacts. § 1.5. PROCESSES OF CHANGE In treating a change which a system undergoes, just by itself or in contact with other systems, thermodynamics confines itself to those changes for which the initial and the final states are thermodynamic states (not neces- sarily thermal equilibrium states). However, the intermediate states through which the system passes may or may not be thermodynamic states. In general, they are very complex states. Only when the process is extremely slow the intermediate states will be thermodynamic, and then the process can be described in terms of the change of a sufficient number of thermodynamic quantities. Cycle: A cycle is a process in which the initial and final states of the system are identical. Infinitesimal process: A change in which the difference between the initial and the final states is infinitesimally small is called an infinitesimal process. Quasi-static process: This is an ideal process during which it may be assumed that both the system and its surroundings maintain thermal equi- librium. Such a process is approximately realized by making the change sufficiently slow. For example, to compress a gas the external pressure p‘) is maintained infinitesimally greater than the gas pressure p, while to expand the gas p is maintained infinitesimally less than p (see Fig. 1.1). In the limit these two processes follow the same path in opposite directions. Thus a quasi-static process is reversible (cf. § 2.1). The following are the important quasi-static processes: i) Quasi-static isothermal process: In this process the system is in contact with an environment of constant temperature (heat bath) and the process occurs quasi-statically at this temperature. ii) Quasi-static adiabatic process: This is a quasi-static process in which the system has no thermal contact (and no mass contact) with its surround- ings but work is done by or on the surroundings. Ch. 1,§ 6] FIRST LAW OF THERMODYNAMICS 5 pressure Teessure | vorumae Fig. 1.1. The quasi-static adiabatic process is sometimes called simply an adiabatic process but since there are irreversible adiabatic processes also it is better to distinguish the two clearly (see § 2.1 for the difference between reversible and irreversible processes). § 1.6. THE FIRST LAW OF THERMODYNAMICS Although the first law of thermodynamics is a particular case of the general law of energy conservation, it implies the existence of the internal energy which is a thermodynamic quantity. THE FIRST LAW OF THERMODYNAMICS. When a system changes from an initial state I to a final state 2, the sum of the work A, the heat Q, and the mass action quantity Z, which it receives from its surroundings, is determined by the states I and 2; it does not depend on the intermediate process. This means that there exists an internal quantity U such that its difference 6 ‘THERMODYNAMIC STATE [ch. 1,86 between the states 1 and 2 is given by U,-U,=A+Q+4+Z. (1.2) The function U/is the internal energy of the system. If the system is character- ised by mechanical energies such as the kinetic energy and the potential energy, eq. (1.2) is replaced by E,-E,=A+Q+Z, (1.2/) where E is the total energy of the system including mechanical energies. The first law of thermodynamics is the law of conservation of energy. It can be expressed as Exoat = constant if E,qai is the energy of the entire system including every subsystem participa- ting in the process of change in question. Eq. (1.2) or eq. (1.2’) is written for a subsystem the energy of which changes as the result of energy-transfer to it from other parts of the system (surroundings) through various mechanisms of contact. The work A, the heat Q, and the mass action Z are respectively the amounts of energy transferred through these contacts. The reader must be familiar with the concepts of work and heat at least from elementary physics. He may be less familiar with that of mass action Z. This represents the amount of energy transferred to the system in a form other than work or heat when a certain mass of matter is transferred from the surroundings. This concept is usually introduced in textbooks of thermo- dynamics much later when heterogeneous systems are treated. But it is included here for the completeness in stating the first law of thermodynamics. Perpetuum mobile of the first kind: If a system undergoes a cycle, A4+Q+Z=0. (13) Thus the system does work, —A, on the surroundings at the expense of Q+Z which it extracts from the surroundings. A perpetuum mobile of the first kind would operate in a cycle which does work on its surroundings without this expense. Therefore, the first law of thermodynamics is also called the principle of the impossibility of a perpetuum mobile of the first kind. Internal energy: The internal energy is the energy within the system, usually excluding the kinetic energy of the system as a whole and the potential energy of the system in an external field (although in some cases the potential energy may be considered as a part of the internal energy). Microscopically, the internal energy is the sum of the molecular kinetic energy and the interaction Ch. 1, § 6] FIRST LAW OF THERMODYNAMICS 7 energy of the material particles, but thermodynamics usually does not ask its origin. The first law of thermodynamics stresses that the internal energy is a state quantity. The additive constant of the internal energy may be fixed by choosing a certain standard state for its zero. Figure 1.2 shows the internal energy of air as a function of temperature and pressure. The point C is the critical point (see example 8). U cal/mol Fig. 1.2, The internal energy of air. The dependence of work, heat and mass action on the process. Unlike U, each of the quantities A, Q and Z in eqs. (1.2) and (1.2’) depends on the process. For example, the same change 12 may be achieved through a mechanical action or a thermal action, or U,-U,;=A (Process I) =Q (Process Il). a4) Hence A, Q and Z are not state quantities nor can they be expressed in terms of the difference of state quantities. The positive statement of the first law is that, nevertheless, the sum A+@Q+Z is determined by the difference in state quantities. 8 ‘THERMODYNAMIC STATE [Ch. 1,§7 DIveRTISSEMENT 1 Founders of the first law of thermodynamics. If a tomb of the Unknown Scientists had been built in the 1850’s, the most appropriate inscription would have been “In memory of the grief and sacrifice of those who fought to realize a perpetuum mobile’’. But the law of conservation of energy, or the first law of thermodynamics, is associated primarily with three great names, Mayer, Helmholtz and Joule. Julius Robert Mayer (1814-1878) was really a genius who was born in this world only with the errand to make this great declaration, Hermann Ludwig Ferdinand von Helmholtz (1821-1894) gave this law the name “rhaltung der Kraft” or “the conservation of energy”. Like Mayer, he started his career as a medical doctor but lived a glorious life as the greatest physiologist and physicist of the day. James Prescott Joule (1818-1889) worked over forty years to establish the experimental verifi- cation of the equivalence of work and heat. Among the three, Mayer was the first who arrived at this law and the last whose work was recognized. His life was most dramatic, A lightening stroke of genius overtook him, a German doctor of the age of twenty six, one day on the sea near Java when he noticed that venous blood of a patient under surgical operation appeared an unusually fresh red. He considered that this might be connected with Lavoisier’s theory of oxi- dation in animals, which process becomes slower in tropical zones be- cause the rate of heat loss by animals will be slower there. A great gener- alization of this observation lead him to the idea of the equivalence of heat and mechanical work. For three years after his voyage, while he was working as a medical doctor at home, he devoted himself to complete the first work on the conservation of energy ““Bemerkungen iiber die Krafte der unbelebten Natur” which was sent to the Poggendorf Annalen and was never published by it. In 1842 Liebig published this paper in his journal (Annalen der Chemie und Pharmacie) but it was ignored for many years. Mayer wrote four papers before 1851. During these years of unusual activity he cared for nothing other than his theory. In 1852 he became mentally deranged and was hospitalized. He recovered after two years but never returned to science. § 1.7. HEAT AND ENTHALPY Ifa change 1-»2 due to thermal action can also occur through a mechanical action (Process I of (1.4), the heat Q which is introduced by the thermal action (Process II) can be measured in terms of the work A. This is the operational definition of heat, and heat in this sense is measured in units of work. The conventional unit is given by 1 cal = 4.1840 joule (thermodynamic calorie*). * Calorie as a practical unit has various definitions. Ch. 1, §8] FIRST LAW OF THERMODYNAMICS 9 The equivalence of heat and work, as indicated by eq. (1.4), means es- sentially the first law. The experimental determination of the heat equivalent of work was achieved by Joule. If the change takes place at constant pressure (isobaric process), the mechanical work associated with the change AV in volume is simply given by —p AV. If there is no other work done, the first law states that U,-U, = p(Y- V2) + Q- Therefore it is convenient to introduce the function H=U+pVv and write H,-H,=Q. (1.4) The quantity H, introduced here, is called the enthalpy, or the heat function. § 1.8. THEFIRSTLAW OF THERMODYNAMICS APPLIED TO INFINITESIMAL PROCESSES For an infinitesimal process, eq. (1.2) can be written as dU =dQ+d'A+d’Z. (5) In particular, if the surroundings consist of a single simple system dA =— pO dV + DX! dx;, (1.6a) dZ = Yu dN; (= y 6 an), (1.6b) jet j=l where p =pressure of the surroundings, V=volume of the system, X{= i-th type force on the system due to the surroundings, x,=the coordinate of the system corresponding to the force X{°, Nj=the number of molecules or the mole number* of the j-th component in the system, and p)=the chemical potential per molecule or mole of the j-th component in the surroundings. In a quasi-static infinitesimal process, the system is almost in equilibrium with its surroundings and d'A =— pdV + DXdx;, (1.7a) d'Z =YnjdN,. (1.7b) 7 * When Nj is the mole number, it is often written as 7 to indicate this and the corre- sponding chemical potential per mole is written as Gy (see § 3.3). Sometimes the mass of component j, Mj, is used instead of the number of molecules or mole number in which case the chemical potential is expressed per gram. In this book the chemical potential is either per molecule or per mole unless otherwise stated. 10 ‘THERMODYNAMIC STATE (Ch. 1,89 In the case of a quasi-static process in general 2 2 2 2 u-u=[eo-frav+y [xex+S{uan,, (1.8) i i 4 1 Hl 1 where p and yp, are the pressure and the chemical potential of the system itself, because they are equal to those of the surroundings, since the system is in equilibrium with the surroundings. Similarly the external forces are in equilibrium with the forces exerted by the system on the surroundings so that the superscript e of X(° may be omitted. The integration in eq. (1.8) is along the path of the process. Notation d’: Since the heat transferred, the mechanical work done and the mass action quantity depend on the process, their infinitesimal values are notexact differentials of state variables. The notation d’ is used to emphasize this point. Work: In general, the force X) which the surroundings exert on the system is defined such that the infinitesimal work done due to a change in the general coordinate x is X) dx. The expression (1.6a) only indicates this definition. In the case of pressure XY —p. Ina quasi-static process, the system and its sur- roundings are in equilibrium and p©) =p; similarly X= X (internal force). Mass interaction: In the same way as above, (1.6b) may be considered as the definition of the chemical potential, u, of the j-th component in the surroundings. Although the meaning of the chemical potential may be no clearer here than that yu‘ characterizes the state of the surroundings with respect to the j-th component, it will be used more extensively starting with Chapter 3. In a quasi-static process p=, § 1.9. TEMPERATURE The temperature is defined as a quantity which prescribes thermal equi- librium between two bodies in thermal contact. If 0, and 0, are the temper- atures of the two bodies, 0, = 0, is the condition for thermal equilibrium, and if 6, >0,, 0, decreases while 0, increases when they are brought into thermal contact. Obviously, when 6,;=0, and 0,=63, then 0,=63, satisfying con- dition (1.1), Conversely, the concept of temperature is based on (1.1) (c.f. problem 23). Although the choice of a temperature scale at this point is arbitrary, its absolute choice will be established with the introduction of the second law of thermodynamics. Anticipating this, the absolute temperature scale or the centigrade temperature scale will be used hereafter. Ch. 1, § 10) HEAT CAPACITY, SPECIFIC HEAT ll The absolute temperature will be denoted by 7. As an empirical temper- ature scale it is possible to use the gas thermometer. An arbitrary temperature scale can be converted into T (°K, Kelvin) by T Vv) FL tim Ve : To +0 (PV oo (cf. Chapter 2, problem 37). The centigrade temperature scale is defined by 0°C =273.15°K. Trripte point of water = 273.16 °K (1.9) § 1.10. HEAT CAPACITY, SPECIFIC HEAT If a quantity of heat d’Q is introduced quasi-statically into a system by a process which, maintaining a quantity x constant, raises the temperature by dT, the quantity C,, defined by dQ ar’ is the heat capacity of the system for this process, The heat capacity per unit mass of the material is its specific heat. The heat capacity for a mole is called the molar heat ot molecular heat. Since d’Q depends on the process it becomes a quantity of state only when the process (or when x) is specified. Examples: heat capacity at constant pressure (isobaric heat capacity), C,, specific heat at constant pressure (isobaric specific heat), c,, heat capacity at constant volume (isovolumic or isochoric heat capacity), Cy, specific heat at constant volume (isovolumic specific heat), cy. It is possible to formally define C,s (=0) for an adiabatic process and Cian, (=00) for an isothermal processes. G= (1.10) DIVERTISSEMENT 2 Why do we have winter heating? The layman will answer: “To make the room warmer.” The student of thermodynamics will perhaps so express it: “To import the lacking (inner, thermal) energy.” If so, then the layman’s answer is right, the scientist’s wrong. We suppose, to correspond to the actual state of affairs, that the pres- sure of the air in a room always equals that of the external air. In the usual notation, the (inner, thermal) energy is, per unit mass, =o. (An additive constant may be neglected.) Then the energy content is, per unit of volume, w= cpT, 12 THERMODYNAMIC STATE [Ch.1,§11 or, taking into account the equation of state, Pip = RT, we have a = ep/R. For air at atmospheric pressure, ut = 0,0604 cal/em’. The energy content of the room is thus independent of the temperature, solely determined by the state of the barometer. The whole of the energy imported by the heating escapes through the pores of the walls of the room to the outside air. I fetch a bottle of claret from the cold cellar and put it to be tempered. in the warm room. It becomes warmer, but the increased energy content is not borrowed from the air of the room but is brought in from outside. Then why do we have heating? For the same reason that life on the earth needs the radiation of the sun. But this does not exist on the incident energy, for the latter apart from a negligible amount is re-radiated, just as aman, in spite of continual absorption of nourishment, maintains a con- stant body-weight. Our conditions of existence require a determinate degree of temperature, and for the maintenance of this there is needed not addition of energy but addition of entropy. Asa student, I read with advantage a small book by F. Wald entitled “The Mistress of the World and her Shadow’’. These meant energy and entropy. In the course of advancing knowledge the two seem to me to have exchanged places. In the huge manufactory of natural processes, the principle of entropy occupies the position of manager, for it dictates the manner and method of the whole business, whilst the principle of energy merely does the bookkeeping, balancing credits and debits. R. EMDEN Kempterstrasse 5, Zurich. The above is a note published in Nature 141 (1938) 908. A. Sommerfeld found it so interesting that he cited it in his book Thermodynamik und Statistik (Vorlesungen iiber theoretische Physik, Bd. 5, Dietrich’sche Verlag, Wiesbaden; English translation by F. Kestin, Academic Press Inc., New York, 1956). R. Emden is known by his work in astrophysics and meteorology as represented by an article in der Enzyklopidie der mathematischen Wissenschaften Thermodynamik der Himmelskorper (Teubner, Leipzig-Berlin, 1926). § 1.11. EQUATION OF STATE The existence of an equation of state: The thermal equilibrium state of a system is usually defined uniquely by specifying the values of a few state variables. For instance, the thermal equilibrium state of a fluid such as a Ch. 1,§ 11] EQUATION OF STATE 13 simple gas or liquid is determined by only two variables, the pressure p and the specific volume v (sometimes we may have more variables; for instance, if the fluid is put in an external electric field or in a magnetic field, the field quantities, E or H, are also variables on which the state of the fluid may depend). This fact, combined with the zeroth law of thermodynamics (1.1), shows that a functional relationship exists between the temperature 8 (on an arbitrarily scale), and the state variables (see problem 23). If the latter are p and », such a relation can be written as 9=f(p,»), which is the equation of state. Similarly it may be shown, for a more general system, that there exists a functional relationship between temperature and a necessary and sufficient number of variables (other than temperature), i.e. that there exists an equation of state. However, thermodynamics does not say anything about its functional form except that it must satisfy certain stability conditions (see § 3.8). Equations of state are determined empirically or for a given system theoretically by statistical mechanics, so that in the framework of thermodynamics they are regarded as something given in order to specify the system. The equation of state of an ideal gas: The general form is pV=g(6), but, if the temperature scale of the gas thermometer is used (it can be shown by the second law of thermodynamics that this agrees with the absolute temper- ature scale, see § 2.5), it is written as pV =nRT = NIT, (1.11) where n is the mole number, N is the number of molecules, R the gas constant and k is the Boltzmann constant: R = 8.314 x 107 erg mol~! deg~! = 1.9872 cal mol”! deg™*, k = 1.3803 x 107'° erg deg = 8.616 x 10°75 eV deg™!. Any real gas becomes ideal in the limit of high temperatures or low densities. It can be shown from the second law of thermodynamics that the internal energy of an ideal gas is independent of the volume, V (cf. Chapter 2, problem 2): (QU/eV);=0, U=U(T). (1.12a) The form of U(T) is determined empirically or theoretically by statistical mechanics. In the normal range of temperature U =nCiT (1.12b) 14 ‘THERMODYNAMIC STATE [Ch.1,§11 2000 Vv 1/mol Fig. 1.3. Pressure of an ideal gas. is a good approximation for most pure gases, so that the specific heat is constant. For those temperatures, the isovolumic molar specific heat, C7. is C}=4R — (monatomic gas) =45R (diatomic gas) =3R (polyatomic gas). (1.13) Mixture of ideal gases: A mixture of r species of ideal gases, consisting of n, moles of the i-th gas (i= 1, 2,..., 7), obeys the equation of state pV =nRT, = n=Yn. (1.14) Corresponding to (1.12), the equation U=YnC2,T (1.15) usually holds for the normal range of temperatures. The partial pressure p, of the i-th component is defined by pW =nRT (1.16) Ch. 1, § 11] EQUATION OF STATE 15 and therefore p=Dp- The ideal gas law applies to real gases as a limiting law: it is a good approximation when the pressure is sufficiently low and the temperature sufficiently high. For higher pressures (higher densities) and lower temper- atures, deviations from the ideal gas law become more noticeable, Figures 1.4 and 1.5 illustrate this for argon. In fact even liquefaction starts when a dense gas is cooled down. There are a number of equations of states to describe dense gases, of which we mention here only the following two. The Van der Waals equation of state is given by (r+ S)e-H=Rr, (1.17) where v= V/n is the molar volume of Table 1.1 the gas. Because of its simple analyt- rr rs ical form, it is often used as an | (atmem®/mol*) | (cm$/mol) approximation for real gases. The || constants a and b for various gases He | 0.03415 x 10° 23.71 are given in Table 1.1. ome lecgaehcicel eer Virial expansion (Kamerlingh-On- A | 1301. x 10°) (30.22 nes’ equation of state): A practical Ne | 1.346 x 108 38.52 Os | 1.361 x 108 | 32.58 expansion for real gaseshasthe form 02/136), x 108 | 32.88 A, As COz | 3.959 x 108 | 42.69 pow rr ft + +44} NO 3.788 x 108 | 44.18 vo HO | 5.468 x10 30.52 (1.18) Cle | 6.501 x 10° 56.26 oe SOz | 6.707 x 108 $6.39 po=RT {1 + Byp + Byp? +--}, (®. S. Epstein, Textbook of thermo- dynamics, John Wiley & Sons, New (1.19) York 1937) where Ap, A3,..., of By, Bs,... ate functions of temperature and are called virial coefficients. Ideal paramagnet and the Curie law: The magnetization M of a magnetic body is a function of the external field H, and the temperature 7. If the equation of state is simply M=f(HIT), (f(0)=0) (1.20) the material is said to be an ideal paramagnet. In particular, when H/T is 16 ‘THERMODYNAMIC STATE [Ch. 1,§12 Argon P= 48.00 atm Ve=1.8839em'/g, Te= 150.72 °K v=P/Pe Fig. 1.4. A relief of the equation of state of argon. small, eq. (1.20) is approximated by M=CH/T (1.21) which is the Curie law, C being the Curie constant. Non-metallic crystals containing paramagnetic metal ions often obey this law. However, this ideal law is bound to break down as the temperature approaches the absolute zero (see Chapter 3, problem 26). § 1.12. CHANGE OF INDEPENDENT VARIABLES Sometimes, in a thermodynamic treatment, p and V are considered to be the independent variables and at other times T and V are so considered. Hence, just writing 0U/@V does not indicate which variables are chosen as independent. To clarify this, the other variables which are kept constant in partial differentiation are indicated by subscripts, e.g., (GU/aV),, (BU/AV)p. Ch. 1, § 12] (CHANGE OF INDEPENDENT VARIABLES 17 Argon Pc =48.00 atm, Ve =1.8839em!/ g, Te =150.72 °K a PV/pVe po. \/ ea a D=P/Pe Fig. 1.5. A relief of the pV-surface of argon. Some of the thermodynamic relations are simply relations between partial derivatives with different choices of independent variables, and are obtained by a change of independent variables. The following identity is particularly useful: when three variables x, y and z have a functional dependence, it is possible to consider z as a function of x and y, or x as a function of y and z, or y asa function of z and x. Then we have a2\ (ax\ (ay\ _ (2)G)().--" om 18 ‘THERMODYNAMIC STATE [Ch.1 (5). ~~ 5/8). (1.23) This follows immediately from a; «-(F) ax+(®) dy. (1.24) ax), ay), If z=const., dz=0 and the ratio dx:dy=(dx/dy), must be (1:23). or Examples A magnetic body is inserted in a coil and is magnetized by the magnetic field created by a current through the coil. For the sake of simplicity, assume that the field, H, and the magnetization, M, are uniform throughout the magnetic body. Show that the work done by the electric source in the process of magnetization is M 4= [iam ° per unit volume of the magnetic body. Assume that the magnetic body has no deformation by magnetization. SoLUTION Let us consider the case where a coil is wound around a sufficiently long cylindrical magnetic body. The magnetic field in a coil with n turns per unit length is H, =4nnJ/c when a current J is passed through the coil. The field H inside the magnetic material is the sum of H, and the demagnetizing field Hy, which, however, is small if the specimen is a long needle. Therefore we may assume here H=H,. If the coil is wound over the entire length of the cylinder, whose length is / and cross-sectional area is o, the flux through the coil is ®=onlB= VnB, where V=al is the volume of the magnetic body and B=H+4nM is the flux density in the magnetic body. When the current J is increased, then H and M, and accordingly B, also change so that a counter- electromotive force —(1/c) db/dt= — V (n/c) 4B/dt is induced, according to Faraday’s induction law. The work which the electric source must do against this counter-electromotive force in a time interval dr as a charge J dt = 4cH, dt/nn n Ch, 1) EXAMPLES 19 passes through each section of the wire, is (1/e) (d@/dt) J dt = 4VH dB/x = 4V(H dH + 4nH dM)/x = d[§VH?/n] +VHdM. Of this work, d[4VH?/z] is the work which is needed to create a field H irrespective of the presence of the magnetic body. Thus, the work H dM can be identified with the work per unit volume required to increase the magneti- zation, M, by dM. Hence, A= {dM is the work per unit volume required Fig. 1.6 to increase the magnetization from 0 to M. (If it were possible to fix the final value of magnetization at M, then the work 4VH?/n would be fed back to the electric source in the process of decreasing H to zero, because then we would have dB=dH.) Let d’g be the heat necessary to change the temperature of a gram of a material by d7 keeping the state quantity x constant. For the sake of simplicity, assume that there are only two independent variables, e.g. the specific volume, v, and the temperature, 7. Show that the specific heat c, is given by the equation dq ws) 4S) gh (2 e,= (22) =(% ee 2 **\ar), \ar),* Wavy * ?f\ar), where w is the internal energy per unit mass and p is the pressure. SoLuTION Since the mass of the system is constant, we need only consider the work done by the pressure. From eqs. (1.5) and (1.7a), the first law of thermodynamics applied to unit mass is d’g=du+p dv. The change du in u due to changes 20 ‘THERMODYNAMIC STATE [Ch. 1 dT and do in T and v is du=(du/éT), d7'+ (du/6v)7 dv. Substituting this into the equation for d’g, we have aa (2) ar+{(%) safer ® Although the parameter x is given as a function of T and v for equilibrium states, it is possible to consider v as a function of T and x taking T and x as the independent variables. Thus if x is kept constant and T is changed, dv= (6v/dT), dT. Substituting this into (1) and dividing by dT we obtain the ans- wer. Show that there is a relation C,=Cy+R (Mayer’s relation) between the isobaric and the isovolumic specific heats per mole of an ideal gas. (Js (4) +R SOLUTION ah \Ih The first law of thermodynamics for an infinitesimal quasi-static process of a mole of gas is d’'Q=dU + pav (1.5), (1.7a)). Since the internal energy U of an ideal gas does not depend on the volume, V, we have dU'=Cy(T) d7 and d'Q=C,(T) dT + pav. qa The isovolumic specific heat (d’Q/dT), is in fact Cy. We réwrite eq. (1) as d'Q =Cy dT + d(pV)—Vdp Q) and substitute pV= RT to obtain dQ =(C, +R)dT—V ap, (69) or Cy = (A'Q/AT)apuo = Cy +R. . Given the density of air at NIP (normal temperature and pressure, that is, 0°C, 1 atm), p=0.00129 g/cm, the specific heat at constant pressure, ¢,=0.238 cal/g deg, and its ratio to the isovolumic specific heat, y=¢,/¢y = 1.41, calculate the work equivalent of heat J. Assume air to be an ideal gas with a volume of 22.41 at NTP. wn Ch. 1] EXAMPLES: 21 SOLUTION The gas constant, R, may be obtained in units of joule/mol deg from the ideal gas equation of state pV=RT. It can also be obtained in units of cal/ mol deg from Mayer’s equation C,—C,=R (example 3) and if we express that by R’, then J=R/R’. The normal state (NTP) is an equilibrium state where the temperature, T=0°C=273°K, and the pressure, p=1 atm= 1.013 x 106 dyne/cm?, The volume of a mole at NTP is V=22.4x 10° cm. Hence: R= pV/T = 1.013 x 10° x 22.4 x 10°/273 = 8.32 x 10” erg/mol deg = 8,32 joule/mol deg. Since the isobaric specific heat per unit mass is the isobaric heat capacity per mole divided by the mass of a mole of air (average molecular weight), m, which is m = pV = 0.00129 x 22.4 x 10° = 28.9 g/mol, we have C, = me, = 28.9 x 0.238 = 6.88 cal/mol deg. The isovolumic specific heat per mole Cy =mcy =me,/ = C,p/y =6.88/1.41 = 4.88 cal/mol deg. Therefore we have R’ = C, — Cy = 2.00 cal/mol deg and J = R/R’ = 8.32/2.00 = 4.16 joule/cal. . (i) Calculate the heat necessary to heat air which has a volume of 27m® at ‘one atmosphere from 0 °C to 20°C at constant volume. (ii) How much heat is necessary if the initial volume at 0 °C is 27m* and the air is heated to 20 °C at constant pressure? (iii) If the air is in an insulated room of 27m> with a small leaking hole connected to the outside at 1 atm, how much heat must be put into the room to raise the temperature slowly from 0 °C to 20 °C? Refer to example 4 for the characteristics of air. Consider the specific heat to be constant. SOLUTION (i) This is a case of isovolumic heating. The mass M of air which has a volume of 27m? at 0 °C and 1 atm is calculated from its density, p=0.00129 g/cm, as M =0.00129 x 27 x 10° = 3.48 x 10*g. 2 THERMODYNAMIC STATE {Ch. 1 The isovolumic heat capacity C; is calculated from the isovolumic specific heat cy = ¢p/y = 0.238/1.41 = 0.169 cal/g deg, Cy = Mey =f0.169 x 3.48 x 10* = 5.88 x 10° cal/deg. Considering the heat capacity as constant, the heat Q,, necessary to raise the temperature from T, to T, is T Oy -[oar-(% — T,) Cy = 20 x 5.88 x 10° = 1.176 x 10° cal. Tr (ii) In the case of isobaric heating we use the isobaric heat capacity C, instead of Cy: C, = Me, = yCy = 1.41 x 5.88 x 10° = 8.29 x 10° cal/deg. Then the desired heat Q, is Q, = (Tz — T:) C, = 20 x 8.29 x 10° = 1.658 x 10° cal. (iii) We may assume that the pressure in the room is maintained at 1 atm since the heating is slow. When the air is heated at constant pressure, p, and constant volume, V, the mass of the air, M(T), in the room at temperature T can be determined from the equation of state pV = RTM/m (m is the average molecular weight of air), or M(T)T'=const. If the mass of the air in the room is M, when the temperature is T,, then M(T)=M,T,/T. Since the air of mass M(T) is being heated at constant pressure, its heat capacity is M(T)c,, so that the heat required to raise the temperature to T is T 1 dT T, Q= | M(T)c,4T =c,M,T, | = = ¢yM,T, log. T T, i Em Since T, =0 °C=273 °K, T;=20 °C=293 °K and C,=8.29 x 10° cal/deg as calculated above, so 293 <= = 2.26 x 10° x 0.0706 = 1,596 x 10° cal. 273 Q = 8.29 x 10° x 273 x log Show that the relation pV’=const. (Poisson’s equation) holds in a quasi- static adiabatic process of an ideal gas and derive the work W which the gas does to its surroundings in a quasi-static adiabatic process from (p,, Vi, T;) to (p2, V2, Tp). Assume the specific heat to be constant. _ Ch. 1] EXAMPLES 23 SOLUTION The equation for the first law in the case of a quasi-static process of an ideal gas is d‘Q=C, dT+p dV (eq. (1) of example 3). Substituting the equation of state, the adiabatic condition d’Q=0 can be written as Cy aT + nRT(dV/V) =0. a Using Mayer’s relation (example 3), C,=C,+mR (for n moles) and inte- grating eq. (1) we have C, logT+(C,—C;) log V=const., which, as C,/C, =y, becomes TV’~! = const. (2a) and, if we use the equation of state, pV? = const. (2b) (We can also get dp/p+dV/V=dT/T from pV=nRT and, eliminating d7/T with eq. (1), dp/p-+ydV/V=0 which gives (2b).) When the volume is changed from V, to V> along the adiabatic line, the work, W, done by the gas is , av y ry W=-A=|pdV= Pity (as pV’ = pV?) Wi _ _ 1 =p VE) = (9M ~ aba) 8) nR (T,-%)=C,(T,-T) (as Cy(y- 1) = C, — Cy = mR). -1 In the case of an adiabatic expansion (V,>V,), T; >>. The gas does work while the internal energy decreases, so that the temperature naturally decreases (cf. Chapter 3, example 5). Show that the following relation exists between the adiabatic compressibility Kya= —(1/v) (6v/Gp),4 and the isothermal compressibility x;= —(1/e) @v/ép)r: Co Kua = — Krys &p where c, and ¢, are the specific heats at constant volume and constant pressure respectively.

You might also like