comparison org/stat-mech.org @ 3:8f3b6dcb9add

Transcribed up to section 1.9, Entropy of an Ideal Boltzmann Gas
author Dylan Holmes <ocsenave@gmail.com>
date Sun, 29 Apr 2012 02:38:22 -0500
parents afbe1fe19b36
children 299a098a30da
comparison
equal deleted inserted replaced
2:afbe1fe19b36 3:8f3b6dcb9add
89 (probability). Our later development of probability theory in 89 (probability). Our later development of probability theory in
90 Chapter 6,7 will be, to a considerable degree, a paraphrase 90 Chapter 6,7 will be, to a considerable degree, a paraphrase
91 of our present review of the logic underlying classical 91 of our present review of the logic underlying classical
92 thermodynamics. 92 thermodynamics.
93 93
94 ** The Primitive Thermometer. 94 ** The Primitive Thermometer
95 95
96 The earliest stages of our 96 The earliest stages of our
97 story are necessarily speculative, since they took place long 97 story are necessarily speculative, since they took place long
98 before the beginnings of recorded history. But we can hardly 98 before the beginnings of recorded history. But we can hardly
99 doubt that primitive man learned quickly that objects exposed 99 doubt that primitive man learned quickly that objects exposed
143 is not specified), any monotonic increasing function 143 is not specified), any monotonic increasing function
144 \(t‘ = f(t)\) provides an equally good temperature scale for the 144 \(t‘ = f(t)\) provides an equally good temperature scale for the
145 present. 145 present.
146 146
147 147
148 ** Thermodynamic Systems. 148 ** Thermodynamic Systems
149 149
150 The \ldquo{}thermodynamic systems\rdquo{} which 150 The \ldquo{}thermodynamic systems\rdquo{} which
151 are the objects of our study may be, physically, almost any 151 are the objects of our study may be, physically, almost any
152 collections of objects. The traditional simplest system with 152 collections of objects. The traditional simplest system with
153 which to begin a study of thermodynamics is a volume of gas. 153 which to begin a study of thermodynamics is a volume of gas.
224 special status, these relations have become known as the 224 special status, these relations have become known as the
225 \ldquo{}laws\rdquo{} 225 \ldquo{}laws\rdquo{}
226 of thermodynamics . The most fundamental one is a qualitative 226 of thermodynamics . The most fundamental one is a qualitative
227 rather than quantitative relation, the \ldquo{}zero'th law.\rdquo{} 227 rather than quantitative relation, the \ldquo{}zero'th law.\rdquo{}
228 228
229 ** Equilibrium; the \ldquo{}Zero‘th Law.\rdquo{} 229 ** Equilibrium; the Zeroth Law
230 230
231 It is a common experience 231 It is a common experience
232 that when objects are placed in contact with each other but 232 that when objects are placed in contact with each other but
233 isolated from their surroundings, they may undergo observable 233 isolated from their surroundings, they may undergo observable
234 changes for a time as a result; one body may become warmer, 234 changes for a time as a result; one body may become warmer,
388 parameters with one constraint) are said to possess two 388 parameters with one constraint) are said to possess two
389 /degrees of freedom/; for the range of possible equilibrium states is defined 389 /degrees of freedom/; for the range of possible equilibrium states is defined
390 by specifying any two of the variables arbitrarily, whereupon the 390 by specifying any two of the variables arbitrarily, whereupon the
391 third, and all others we may introduce, are determined. 391 third, and all others we may introduce, are determined.
392 Mathematically, this is expressed by the existence of a functional 392 Mathematically, this is expressed by the existence of a functional
393 relationship of the form[fn:: /Edit./: The set of solutions to an equation 393 relationship of the form[fn:: The set of solutions to an equation
394 like /f(X,x,t)=/ const. is called a /level set/. Here, Jaynes is 394 like /f(X,x,t)=/ const. is called a /level set/. Here, Jaynes is
395 saying that the quantities /X/, /x/, and /t/ follow a \ldquo{}functional 395 saying that the quantities /X/, /x/, and /t/ follow a \ldquo{}functional
396 rule\rdquo{}, so the set of physically allowed combinations of /X/, 396 rule\rdquo{}, so the set of physically allowed combinations of /X/,
397 /x/, and /t/ in equilibrium states can be 397 /x/, and /t/ in equilibrium states can be
398 expressed as the level set of a function. 398 expressed as the level set of a function.
751 \begin{equation} 751 \begin{equation}
752 \sum_{j=0}^n K_j M_j \Delta t_j = 0 752 \sum_{j=0}^n K_j M_j \Delta t_j = 0
753 \end{equation} 753 \end{equation}
754 is always satisfied. This sort of process is an old story in 754 is always satisfied. This sort of process is an old story in
755 scientific investigations; although the great theoretician Boltzmann 755 scientific investigations; although the great theoretician Boltzmann
756 is said to have remarked: \ldquo{}Elegance is for tailors \rdquo{}, it 756 is said to have remarked: \ldquo{}Elegance is for tailors\rdquo{}, it
757 remains true that the attempt to reduce equations to the most 757 remains true that the attempt to reduce equations to the most
758 symmetrical form has often suggested important generalizations of 758 symmetrical form has often suggested important generalizations of
759 physical laws, and is a great aid to memory. Witness Maxwell's 759 physical laws, and is a great aid to memory. Witness Maxwell's
760 \ldquo{}displacement current\rdquo{}, which was needed to fill in a 760 \ldquo{}displacement current\rdquo{}, which was needed to fill in a
761 gap and restore the symmetry of the electromagnetic equations; as soon 761 gap and restore the symmetry of the electromagnetic equations; as soon
765 for we recognize that (1-12) has the standard form of a /conservation 765 for we recognize that (1-12) has the standard form of a /conservation
766 law/; it defines a new quantity which is conserved in thermal 766 law/; it defines a new quantity which is conserved in thermal
767 interactions of the type just studied. 767 interactions of the type just studied.
768 768
769 The similarity of (1-12) to conservation laws in general may be seen 769 The similarity of (1-12) to conservation laws in general may be seen
770 as follows. Let $A$ be some quantity that is conserved; the $i$th 770 as follows. Let $A$ be some quantity that is conserved; the \(i\)th
771 system has an amount of it $A_i$. Now when the systems interact such 771 system has an amount of it $A_i$. Now when the systems interact such
772 that some $A$ is transferred between them, the amount of $A$ in the 772 that some $A$ is transferred between them, the amount of $A$ in the
773 $i$th system is changed by a net amount \(\Delta A_i = (A_i)_{final} - 773 \(i\)th system is changed by a net amount \(\Delta A_i = (A_i)_{final} -
774 (A_i)_{initial}\); and the fact that there is no net change in the 774 (A_i)_{initial}\); and the fact that there is no net change in the
775 total amount of $A$ is expressed by the equation \(\sum_i \Delta 775 total amount of $A$ is expressed by the equation \(\sum_i \Delta
776 A_i = 0$. Thus, the law of conservation of matter in a chemical 776 A_i = 0\). Thus, the law of conservation of matter in a chemical
777 reaction is expressed by \(\sum_i \Delta M_i = 0\), where $M_i$ is the 777 reaction is expressed by \(\sum_i \Delta M_i = 0\), where $M_i$ is the
778 mass of the $i$th chemical component. 778 mass of the \(i\)th chemical component.
779 779
780 what is this new conserved quantity? Mathematically, it can be defined 780 What is this new conserved quantity? Mathematically, it can be defined
781 as $Q_i = K_i\cdot M_i cdot t_i; whereupon (1-12) becomes 781 as $Q_i = K_i\cdot M_i \cdot t_i$; whereupon (1-12) becomes
782 782
783 \begin{equation} 783 \begin{equation}
784 \sum_i \Delta Q_i = 0 784 \sum_i \Delta Q_i = 0
785 \end{equation} 785 \end{equation}
786 786
881 continue to teach it and to use it until we have something better to 881 continue to teach it and to use it until we have something better to
882 put in its place.) 882 put in its place.)
883 883
884 # what is "the specific heat of a gas at constant pressure/volume"? 884 # what is "the specific heat of a gas at constant pressure/volume"?
885 # changed t for temperature below from capital T to lowercase t. 885 # changed t for temperature below from capital T to lowercase t.
886 Another failure of the conservation law (1-13) was noted in 1842 by 886 Another failure of the conservation law (1-13) was [[http://web.lemoyne.edu/~giunta/mayer.html][noted in 1842]] by
887 R. Mayer, a German physician, who pointed out that the data already 887 R. Mayer, a German physician, who pointed out that the data already
888 available showed that the specific heat of a gas at constant pressure, 888 available showed that the specific heat of a gas at constant pressure,
889 C_p, was greater than at constant volume $C_v$. He surmised that the 889 C_p, was greater than at constant volume $C_v$. He surmised that the
890 difference was due to the work done in expansion of the gas against 890 difference was due to the work done in expansion of the gas against
891 atmospheric pressure, when measuring $C_p$. Supposing that the 891 atmospheric pressure, when measuring $C_p$. Supposing that the
903 conservation law (1-13) exists whenever purely thermal interactions 903 conservation law (1-13) exists whenever purely thermal interactions
904 were involved; but in processes involving mechanical work, the 904 were involved; but in processes involving mechanical work, the
905 conservation law broke down. 905 conservation law broke down.
906 906
907 ** The First Law 907 ** The First Law
908 Corresponding to the partially valid law of \ldquo{}conservation of
909 heat\rdquo{}, there had long been known another partially valid
910 conservation law in mechanics. The principle of conservation of
911 mechanical energy had been given by Leibnitz in 1693 in noting that,
912 according to the laws of Newtonian mechanics, one could define
913 potential and kinetic energy so that in mechanical processes they were
914 interconverted into each other, the total energy remaining
915 constant. But this too was not universally valid---the mechanical
916 energy was conserved only in the absence of frictional forces. In
917 processes involving friction, the mechanical energy seemed to
918 disappear.
919
920 So we had a law of conservation of heat, which broke down whenever
921 mechanical work was done; and a law of conservation of mechanical
922 energy, which broke down when frictional forces were present. If, as
923 Mayer had suggested, heat was itself a form of energy, then one had
924 the possibility of accounting for both of these failures in a new law
925 of conservation of /total/ (mechanical + heat) energy. On one hand,
926 the difference $C_p-C_v$ of heat capacities of gases would be
927 accounted for by the mechanical work done in expansion; on the other
928 hand, the disappearance of mechanical energy would be accounted for by
929 the heat produced by friction.
930
931 But to establish this requires more than just suggesting the idea and
932 illustrating its application in one or two cases --- if this is really
933 a new conservation law adequate to replace the two old ones, it must
934 be shown to be valid for /all/ substances and /all/ kinds of
935 interaction. For example, if one calorie of heat corresponded to $E$
936 ergs of mechanical energy in the gas experiments, but to a different
937 amoun $E^\prime$ in heat produced by friction, then there would be no
938 universal conservation law. This \ldquo{}first law\rdquo{} of
939 thermodynamics must therefore take the form:
940 #+begin_quote
941 There exists a /universal/ mechanical equivalent of heat, so that the
942 total (mechanical energy) + (heat energy) remeains constant in all
943 physical processes.
944 #+end_quote
945
946 It was James Prescott Joule who provided the [[http://www.chemteam.info/Chem-History/Joule-Heat-1845.html][first experimental data]]
947 indicating this universality, and providing the first accurate
948 numerical value of this mechanical equivalent. The calorie had been
949 defined as the amount of heat required to raise the temperature of one
950 gram of water by one degree Centigrade (more precisely, to raise it
951 from 14.5 to 15.5$^\circ C$). Joule measured the heating of a number
952 of different liquids due to mechanical stirring and electrical
953 heating, and established that, within the experimental accuracy (about
954 one percent) a /calorie/ of heat always corresponded to the same
955 amount of energy. Modern measurements give this numerical value as: 1
956 calorie = 4.184 \times 10^7 ergs = 4.184 joules.
957 # capitalize Joules? I think the convention is to spell them out in lowercase.
958
959 The circumstances of this important work are worth noting. Joule was
960 in frail health as a child, and was educated by private tutors,
961 including the chemist, John Dalton, who had formulated the atomic
962 hypothesis in the early nineteenth century. In 1839, when Joule was
963 nineteen, his father (a wealthy brewer) built a private laboratory for
964 him in Manchester, England; and the good use he made of it is shown by
965 the fact that, within a few months of the opening of this laboratory
966 (1840), he had completed his first important piece of work, at the
967 age of twenty. This was his establishment of the law of \ldquo{}Joule
968 heating,\rdquo{} $P=I^2 R$, due to the electric current in a
969 resistor. He then used this effect to determine the universality and
970 numerical value of the mechanical equivalent of heat, reported
971 in 1843. His mechanical stirring experiments reported in 1849 yielded
972 the value 1 calorie = 4.154 \times 10^7 ergs, amount 0.7% too low;
973 this determination was not improved upon for several decades.
974
975 The first law of thermodynamics may then be stated mathematically as
976 follows:
977
978 #+begin_quote
979 There exists a state function (i.e., a definite function of the
980 thermodynamic state) $U$, representing the total energy of any system,
981 such that in any process in which we change from one equilibrium to
982 another, the net change in $U$ is given by the difference of the heat
983 $Q$ supplied to the system, and the mechanical work $W$ done by the
984 system.
985 #+end_quote
986 On an infinitesimal change of state, this becomes
987
988 \begin{equation}
989 dU = dQ - dW.
990 \end{equation}
991
992 For a system of two degrees of freedom, defined by pressure $P$,
993 volume $V$, and temperature $t$, we have $dW = PdV$. Then if we regard
994 $U$ as a function $U(V,t)$ of volume and temperature, the fact that
995 $U$ is a state function means that $dU$ must be an exact differential;
996 i.e., the integral
997
998 \begin{equation}
999 \int_1^2 dU = U(V_2,t_2) - U(V_1,t_1)
1000 \end{equation}
1001 between any two thermodynamic states must be independent of the
1002 path. Equivalently, the integral $\oint dU$ over any closed cyclic
1003 path (for example, integrate from state 1 to state 2 along path A,
1004 then back to state 1 by a different path B) must be zero. From (1-15),
1005 this gives for any cyclic integral,
1006
1007 \begin{equation}
1008 \oint dQ = \oint P dV
1009 \end{equation}
1010
1011 another form of the first law, which states that in any process in
1012 which the system ends in the same thermodynamic state as the initial
1013 one, the total heat absorbed by the system must be equal to the total
1014 work done.
1015
1016 Although the equations (1-15)-(1-17) are rather trivial
1017 mathematically, it is important to avoid later conclusions that we
1018 understand their exact meaning. In the first place, we have to
1019 understand that we are now measuring heat energy and mechanical energy
1020 in the same units; i.e. if we measured $Q$ in calories and $W$ in
1021 ergs, then (1-15) would of course not be correct. It does
1022 not matter whether we apply Joule's mechanical equivalent of heat
1023 to express $Q$ in ergs, or whether we apply it in the opposite way
1024 to express $U$ and $W$ in calories; each procedure will be useful in
1025 various problems. We can develop the general equations of
1026 thermodynamics
1027 without committing ourselves to any particular units,
1028 but of course all terms in a given equation must be expressed
1029 in the same units.
1030
1031 Secondly, we have already stressed that the theory being
1032 developed must, strictly speaking, be a theory only of
1033 equilibrium states, since otherwise we have no operational definition
1034 of temperature . When we integrate over any \ldquo{}path\rdquo{} in the $(V-t)$
1035 plane, therefore, it must be understood that the path of
1036 integration is, strictly speaking, just a /locus of equilibrium
1037 states/; nonequilibrium states cannot be represented by points
1038 in the $(V-t)$ plane.
1039
1040 But then, what is the relation between path of equilibrium
1041 states appearing in our equations, and the sequence of conditions
1042 produced experimentally when we change the state of a system in
1043 the laboratory? With any change of state (heating, compression,
1044 etc.) proceeding at a finite rate we do not have equilibrium in
1045 termediate states; and so there is no corresponding \ldquo{}path\rdquo{} in
1046 the $(V-t)$ plane ; only the initial and final equilibrium states
1047 correspond to definite points. But if we carry out the change
1048 of state more and more slowly, the physical states produced are
1049 nearer and nearer to equilibrium state. Therefore, we interpret
1050 a path of integration in the $(V-t)$ plane, not as representing
1051 the intermediate states of any real experiment carried out at
1052 a finite rate, but as the /limit/ of this sequence of states, in
1053 the limit where the change of state takes place arbitrarily
1054 slowly.
1055
1056 An arbitrarily slow process, so that we remain arbitrarily
1057 near to equilibrium at all times, has another important property.
1058 If heat is flowing at an arbitrarily small rate, the temperature
1059 difference producing it must be arbitrarily small, and therefore
1060 an arbitrarily small temperature change would be able to reverse
1061 the direction of heat flow. If the Volume is changing very
1062 slowly, the pressure difference responsible for it must be very
1063 small; so a small change in pressure would be able to reverse
1064 the direction of motion. In other words, a process carried out
1065 arbitrarily slowly is /reversible/; if a system is arbitrarily
1066 close to equilibrium, then an arbitrarily small change in its
1067 environment can reverse the direction of the process.
1068 Recognizing this, we can then say that the paths of integra
1069 tion in our equations are to be interpreted physically as
1070 /reversible paths/ . In practice, some systems (such as gases)
1071 come to equilibrium so rapidly that rather fast changes of
1072 state (on the time scale of our own perceptions) may be quite
1073 good approximations to reversible changes; thus the change of
1074 state of water vapor in a steam engine may be considered
1075 reversible to a useful engineering approximation.
1076
1077
1078 ** Intensive and Extensive Parameters
1079
1080 The literature of thermodynamics has long recognized a distinction between two
1081 kinds of quantities that may be used to define the thermodynamic
1082 state. If we imagine a given system as composed of smaller
1083 subsystems, we usually find that some of the thermodynamic variables
1084 have the same values in each subsystem, while others are additive,
1085 the total amount being the sum of the values of each subsystem.
1086 These are called /intensive/ and /extensive/ variables, respectively.
1087 According to this definition, evidently, the mass of a system is
1088 always an extensive quantity, and at equilibrium the temperature
1089 is an intensive ‘quantity. Likewise, the energy will be extensive
1090 provided that the interaction energy between the subsystems can
1091 be neglected.
1092
1093 It is important to note, however, that in general the terms
1094 \ldquo{}intensive\rdquo{} and \ldquo{}extensive\rdquo{}
1095 so defined cannot be regarded as
1096 establishing a real physical distinction between the variables.
1097 This distinction is, like the notion of number of degrees of
1098 freedom, in part an anthropomorphic one, because it may depend
1099 on the particular kind of subdivision we choose to imagine. For
1100 example, a volume of air may be imagined to consist of a number
1101 of smaller contiguous volume elements. With this subdivision,
1102 the pressure is the same in all subsystems, and is therefore in
1103 tensive; while the volume is additive and therefore extensive .
1104 But we may equally well regard the volume of air as composed of
1105 its constituent nitrogen and oxygen subsystems (or we could re
1106 gard pure hydrogen as composed of two subsystems, in which the
1107 molecules have odd and even rotational quantum numbers
1108 respectively, etc.) . With this kind of subdivision the volume is the
1109 same in all subsystems, while the pressure is the sum of the
1110 partial pressures of its constituents; and it appears that the
1111 roles of \ldquo{}intensive\rdquo{} and \ldquo{}extensive\rdquo{}
1112 have been interchanged. Note that this ambiguity cannot be removed by requiring
1113 that we consider only spatial subdivisions, such that each sub
1114 system has the same local composi tion . For, consider a s tressed
1115 elastic solid, such as a stretched rubber band. If we imagine
1116 the rubber band as divided, conceptually, into small subsystems
1117 by passing planes through it normal to its axis, then the tension
1118 is the same in all subsystems, while the elongation is additive.
1119 But if the dividing planes are parallel to the axis, the elonga
1120 tion is the same in all subsystems, while the tension is
1121 additive; once again, the roles of \ldquo{}extensive\rdquo{} and
1122 \ldquo{}intensive\rdquo{} are
1123 interchanged merely by imagining a different kind of subdivision.
1124 In spite of the fundamental ambiguity of the usual definitions,
1125 the notions of extensive and intensive variables are useful,
1126 and in practice we seem to have no difficulty in deciding
1127 which quantities should be considered intensive. Perhaps the
1128 distinction is better characterized, not by considering
1129 subdivisions at all, but by adopting a different definition, in which
1130 we recognize that some quantities have the nature of a \ldquo{}force\rdquo{}
1131 or \ldquo{}potential\rdquo{}, or some other local physical property, and are
1132 therefore called intensive, while others have the nature of a
1133 \ldquo{}displacement\rdquo{} or a \ldquo{}quantity\rdquo{} of
1134 something (i.e. are proportional to the size of the system),
1135 and are therefore called extensive. Admittedly, this definition is somewhat vague, in a
1136 way that can also lead to ambiguities ; in any event, let us agree
1137 to class pressure, stress tensor, mass density, energy density,
1138 particle density, temperature, chemical potential, angular
1139 velocity, as intensive, while volume, mass, energy, particle
1140 numbers, strain, entropy, angular momentum, will be considered
1141 extensive.
1142
1143 ** The Kelvin Temperature Scale
1144 The form of the first law,
1145 $dU = dQ - dW$, expresses the net energy increment of a system as
1146 the heat energy supplied to it, minus the work done by it. In
1147 the simplest systems of two degrees of freedom, defined by
1148 pressure and volume as the thermodynamic variables, the work done
1149 in an infinitesimal reversible change of state can be separated
1150 into a product $dW = PdV$ of an intensive and an extensive quantity.
1151 Furthermore, we know that the pressure $P$ is not only the
1152 intensive factor of the work; it is also the \ldquo{}potential\rdquo{}
1153 which governs mechanical equilibrium (in this case, equilibrium with respect
1154 to exchange of volume) between two systems; i .e., if they are
1155 separated by a flexible but impermeable membrane, the two systems
1156 will exchange volume $dV_1 = -dV_2$ in a direction determined by the
1157 pressure difference, until the pressures are equalized. The
1158 energy exchanged in this way between the systems is a product
1159 of the form
1160 #+begin_quote
1161 (/intensity/ of something) \times (/quantity/ of something exchanged)
1162 #+end_quote
1163
1164 Now if heat is merely a particular form of energy that can
1165 also be exchanged between systems, the question arises whether
1166 the quantity of heat energy $dQ$ exchanged in an infinitesimal
1167 reversible change of state can also be written as a product of one
1168 factor which measures the \ldquo{}intensity\rdquo{} of the heat,
1169 times another that represents the \ldquo{}quantity\rdquo{}
1170 of something exchanged between
1171 the systems, such that the intensity factor governs the
1172 conditions of thermal equilibrium and the direction of heat exchange,
1173 in the same way that pressure does for volume exchange.
1174
1175
1176 But we already know that the /temperature/ is the quantity
1177 that governs the heat flow (i.e., heat flows from the hotter to
1178 the cooler body until the temperatures are equalized) . So the
1179 intensive factor in $dQ$ must be essentially the temperature. But
1180 our temperature scale is at present still arbitrary, and we can
1181 hardly expect that such a factorization will be possible for all
1182 calibrations of our thermometers.
1183
1184 The same thing is evidently true of pressure; if instead of
1185 the pressure $P$ as ordinarily defined, we worked with any mono
1186 tonic increasing function $P_1 = P_1 (P)$ we would find that $P_1$ is
1187 just as good as $P$ for determining the direction of volume
1188 exchange and the condition of mechanical equilibrium; but the work
1189 done would not be given by $PdV$; in general, it could not even
1190 be expressed in the form $P_1 \cdot dF(V)$, where $F(V)$ is some function
1191 of V.
1192
1193
1194 Therefore we ask: out of all the monotonic functions $t_1(t)$
1195 corresponding to different empirical temperature scales, is
1196 there one (which we denote as $T(t)$) which forms a \ldquo{}natural\rdquo{}
1197 intensity factor for heat, such that in a reversible change
1198 $dQ = TdS$, where $S(U,V)$ is a new function of the thermodynamic
1199 state? If so, then the temperature scale $T$ will have a great
1200 theoretical advantage, in that the laws of thermodynamics will
1201 take an especially simple form in terms of this particular scale,
1202 and the new quantity $S$, which we call the /entropy/, will be a
1203 kind of \ldquo{}volume\rdquo{} factor for heat.
1204
1205 We recall that $dQ = dU + PdV$ is not an exact differential;
1206 i.e., on a change from one equilibrium state to another the
1207 integral
1208
1209 \[\int_1^2 dQ\]
1210
1211 cannot be set equal to the difference $Q_2 - Q_1$ of values of any
1212 state function $Q(U,V)$, since the integral has different values
1213 for different paths connecting the same initial and final states.
1214 Thus there is no \ldquo{}heat function\rdquo{} $Q(U,V)$, and the notion of
1215 \ldquo{}amount of heat\rdquo{} $Q$ stored in a body has no meaning
1216 (nor does the \ldquo{}amount of work\rdquo{} $W$;
1217 only the total energy is a well-defined quantity).
1218 But we want the entropy $S(U,V)$ to be a definite quantity,
1219 like the energy or volume, and so $dS$ must be an exact differential.
1220 On an infinitesimal reversible change from one equilibrium state
1221 to another, the first law requires that it satisfy[fn:: The first
1222 equality comes from our requirement that $dQ = T\,dS$. The second
1223 equality comes from the fact that $dU = dQ - dW$ (the first law) and
1224 that $dW = PdV$ in the case where the state has two degrees of
1225 freedom, pressure and volume.]
1226
1227 \begin{equation}
1228 dS(U,V) = \frac{dQ}{T} = \frac{dU}{T} + \frac{P}{T}dV
1229 \end{equation}
1230
1231 Thus $(1/T)$ must be an /integrating factor/ which converts $dQ$ into
1232 an exact differential [[fn::A differential $M(x,y)dx +
1233 N(x,y)dy$ is called /exact/ if there is a scalar function
1234 $\Phi(x,y)$ such that $M = \frac{\partial \Phi}{\partial x}$ and
1235 $N=\frac{\partial \Phi}{\partial y}$. If there is, \Phi is called the
1236 /potential function/ of the differential, Conceptually, this means
1237 that M(x,y)dx + N(x,y) dy is the derivative of a scalar potential and
1238 so consequently corresponds to a conservative field.
1239
1240 Even if there is no such potential function
1241 \Phi for the given differential, it is possible to coerce an
1242 inexact differential into an exact one by multiplying by an unknown
1243 function $\mu(x,y)$ (called an /integrating factor/) and requiring the
1244 resulting differential $\mu M\, dx + \mu N\, dy$ to be exact.
1245
1246 To complete the analogy, here we have the differential $dQ =
1247 dU + PdV$ (by the first law) which is not exact---conceptually, there
1248 is no scalar potential nor conserved quantity corresponding to
1249 $dQ$. We have introduced a new differential $dS = \frac{1}{T}dQ$, and we
1250 are searching for the temperature scale $T(U,V)$ which makes $dS$
1251 exact (i.e. which makes $S$ correspond to a conserved quantity). This means
1252 that $\frac{1}{T}$ is playing the role of the integrating factor
1253 \ldquo{}\mu\rdquo{} for the differential $dQ$.]]
1254
1255 Now the question of the existence and properties of
1256 integrating factors is a purely mathematical one, which can be
1257 investigated independently of the properties of any particular
1258 substance. Let us denote this integrating factor for the moment
1259 by $w(U,V) = T^{-1}$; then the first law becomes
1260
1261 \begin{equation}
1262 dS(U,V) = w dU + w P dV
1263 \end{equation}
1264
1265 from which the derivatives are
1266
1267 \begin{equation}
1268 \left(\frac{\partial S}{\partial U}\right)_V = w, \qquad
1269 \left(\frac{\partial S}{\partial V}\right)_U = wP.
1270 \end{equation}
1271
1272 The condition that $dS$ be exact is that the cross-derivatives be
1273 equal, as in (1-4):
1274
1275 \begin{equation}
1276 \frac{\partial^2 S}{\partial U \partial V} = \frac{\partial^2
1277 S}{\partial V \partial U},
1278 \end{equation}
1279
1280 or
1281
1282 \begin{equation}
1283 \left(\frac{\partial w}{\partial V}\right)_U = \left(\frac{\partial
1284 P}{\partial U}\right)_V + P\cdot \left(\frac{\partial w}{\partial U}\right)_V.
1285 \end{equation}
1286
1287 Any function $w(U,V)$ satisfying this differential equation is an
1288 integrating factor for $dQ$.
1289
1290 But if $w(U,V)$ is one such integrating factor, which leads
1291 to the new state function $S(U,V)$, it is evident that
1292 $w_1(U,V) \equiv w \cdot f(S)$ is an equally good integrating factor, where
1293 $f(S)$ is an arbitrary function. Use of $w_1$ will lead to a
1294 different state function
1295
1296 #what's with the variable collision?
1297 \begin{equation}
1298 S_1(U,V) = \int^S f(S) dS
1299 \end{equation}
1300
1301 The mere conversion of into an exact differential is, therefore,
1302 not enough to determine any unique entropy function $S(U,V)$.
1303 However, the derivative
1304
1305 \begin{equation}
1306 \left(\frac{dU}{dV}\right)_S = -P
1307 \end{equation}
1308
1309 is evidently uniquely determined; so also, therefore, is the
1310 family of lines of constant entropy, called /adiabats/, in the
1311 $(U-V)$ plane. But, as (1-24) shows, the numerical value of $S$ on
1312 each adiabat is still completely undetermined.
1313
1314 In order to fix the relative values of $S$ on different
1315 adiabats we need to add the condition, not yet put into the equations,
1316 that the integrating factor $w(U,V) = T^{-1}$ is to define a new
1317 temperature scale . In other words, we now ask: out of the
1318 infinite number of different integrating factors allowed by
1319 the differential equation (1-23), is it possible to find one
1320 which is a function only of the empirical temperature $t$? If
1321 $w=w(t)$, we can write
1322
1323 \begin{equation}
1324 \left(\frac{\partial w}{\partial V}\right)_U = \frac{dw}{dt}\left(\frac{\partial
1325 t}{\partial V}\right)_U
1326 \end{equation}
1327 \begin{equation}
1328 \left(\frac{\partial w}{\partial U}\right)_V = \frac{dw}{dt}\left(\frac{\partial
1329 t}{\partial U}\right)_V
1330 \end{equation}
1331
1332
1333 and (1-23) becomes
1334 \begin{equation}
1335 \frac{d}{dt}\log{w} = \frac{\left(\frac{\partial P}{\partial
1336 U}\right)_V}{\left(\frac{\partial t}{\partial V}\right)_U-P\left(\frac{\partial t}{\partial U}\right)_V}
1337 \end{equation}
1338
1339
1340 which shows that $w$ will be determined to within a multiplicative
1341 factor.
1342
1343 Is the temperature scale thus defined independent of the
1344 empirical scale from which we started? To answer this, let
1345 $t_1 = t_1(t)$ be any monotonic function which defines a different
1346 empirical temperature scale. In place of (1-28), we then have
1347
1348 \begin{equation}
1349 \frac{d}{dt_1}\log{w} \quad=\quad \frac{\left(\frac{\partial P}{\partial
1350 U}\right)_V}{\left(\frac{\partial t_1}{\partial V}\right)_U-P\left(\frac{\partial t_1}{\partial U}\right)_V}
1351 \quad = \quad
1352 \frac{\left(\frac{\partial P}{\partial
1353 U}\right)_V}{\frac{dt_1}{dt}\left[ \left(\frac{\partial t}{\partial
1354 V}\right)_U-P\left(\frac{\partial t}{\partial U}\right)_V\right]},
1355 \end{equation}
1356 or
1357 \begin{equation}
1358 \frac{d}{dt_1}\log{w_1} = \frac{dt}{dt_1}\frac{d}{dt}\log{w}
1359 \end{equation}
1360
1361 which reduces to $d \log{w_1} = d \log{w}$, or
1362 \begin{equation}
1363 w_1 = C\cdot w
1364 \end{equation}
1365
1366 Therefore, integrating factors derived from whatever empirical
1367 temperature scale can differ among themselves only by a
1368 multiplicative factor. For any given substance, therefore, except
1369 for this factor (which corresponds just to our freedom to choose
1370 the size of the units in which we measure temperature), there is
1371 only /one/ temperature scale $T(t) = 1/w$ with the property that
1372 $dS = dQ/T$ is an exact differential.
1373
1374 To find a feasible way of realizing this temperature scale
1375 experimentally, multiply numerator and denominator of the right
1376 hand side of (1-28) by the heat capacity at constant volume,
1377 $C_V^\prime = (\partial U/\partial t) V$, the prime denoting that
1378 it is in terms of the empirical temperature scale $t$.
1379 Integrating between any two states denoted 1 and 2, we have
1380
1381 \begin{equation}
1382 \frac{T_1}{T_2} = \exp\left\{\int_{t_1}^{t_2}
1383 \frac{\left(\frac{\partial P}{\partial t}\right)_V dt}{P - C_V^\prime
1384 \left(\frac{\partial t}{\partial V}\right)_U} \right\}
1385 \end{equation}
1386
1387 If the quantities on the right-hand side have been determined
1388 experimentally, then a numerical integration yields the ratio
1389 of Kelvin temperatures of the two states.
1390
1391 This process is particularly simple if we choose for our
1392 system a volume of gas with the property found in Joule's famous
1393 expansion experiment; when the gas expands freely into a vacuum
1394 (i.e., without doing work, or $U = \text{const.}$), there is no change in
1395 temperature. Real gases when sufficiently far from their condensation
1396 points are found to obey this rule very accurately.
1397 But then
1398
1399 \begin{equation}
1400 \left(\frac{dt}{dV}\right)_U = 0
1401 \end{equation}
1402
1403 and on a change of state in which we heat this gas at constant
1404 volume, (1-31) collapses to
1405
1406 \begin{equation}
1407 \frac{T_1}{T_2} = \exp\left\{\int_{t_1}^{t_2}
1408 \frac{1}{P}\left(\frac{\partial P}{\partial t}\right)_V dt\right\} = \frac{P_2}{P_1}.
1409 \end{equation}
1410
1411 Therefore, with a constant-volume ideal gas thermometer, (or more
1412 generally, a thermometer using any substance obeying (1-32) and
1413 held at constant volume), the measured pressure is directly
1414 proportional to the Kelvin temperature.
1415
1416 For an imperfect gas, if we have measured $(\partial t /\partial
1417 V)_U$ and $C_V^\prime$, Eq. (1-31) determines the necessary
1418 corrections to (1-33). However, an alternative form of (1-31), in
1419 which the roles of pressure and volume are interchanged, proves to be
1420 more convenient for experimental determinations. To derive it, introduce the
1421 enthalpy function
1422
1423 \begin{equation}H = U + PV\end{equation}
1424
1425 with the property
1426
1427 \begin{equation}
1428 dH = dQ + VdP
1429 \end{equation}
1430
1431 Equation (1-19) then becomes
1432
1433 \begin{equation}
1434 dS = \frac{dH}{T} - \frac{V}{T}dP.
1435 \end{equation}
1436
1437 Repeating the steps (1-20) to (1-31) of the above derivation
1438 starting from (1-36) instead of from (1-19), we arrive at
1439
1440 \begin{equation}
1441 \frac{T_2}{T_1} = \exp\left\{\int_{t_1}^{t_2}
1442 \frac{\left(\frac{dV}{dt}\right)_P dt}{V + C_P^\prime
1443 \left(\frac{\partial t}{\partial P}\right)_H}\right\}
1444 \end{equation}
1445
1446 or
1447
1448 \begin{equation}
1449 \frac{T_2}{T_1} = \exp\left\{\frac{\alpha^\prime
1450 dt}{1+\left(C_P^\prime \cdot \mu^\prime / V\right)}\right\}
1451 \end{equation}
1452
1453 where
1454 \begin{equation}
1455 \alpha^\prime \equiv \frac{1}{V}\left(\frac{\partial V}{\partial t}\right)_P
1456 \end{equation}
1457 is the thermal expansion coefficient,
1458 \begin{equation}
1459 C_P^\prime \equiv \left(\frac{\partial H}{\partial t}\right)_P
1460 \end{equation}
1461 is the heat capacity at constant pressure, and
1462 \begin{equation}
1463 \mu^\prime \equiv \left(\frac{dt}{dP}\right)_H
1464 \end{equation}
1465
1466 is the coefficient measured in the Joule-Thompson porous plug
1467 experiment, the primes denoting again that all are to be measured
1468 in terms of the empirical temperature scale $t$.
1469 Since $\alpha^\prime$, $C_P^\prime$, $\mu^\prime$ are all
1470 easily measured in the laboratory, Eq. (1-38) provides a
1471 feasible way of realizing the Kelvin temperature scale experimentally,
1472 taking account of the imperfections of real gases.
1473 For an account of the work of Roebuck and others based on this
1474 relation, see [[http://books.google.com/books?id=KKJKAAAAMAAJ][Zemansky (1943)]]; pp. 252-255.
1475
1476 Note that if $\mu^\prime = O$ and we heat the gas at constant
1477 pressure, (1-38) reduces to
1478
1479 \begin{equation}
1480 \frac{T_2}{T_1} = \exp\left\{ \int_{t_1}^{t_2}
1481 \frac{1}{V}\left(\frac{\partial V}{\partial t}\right)_P dt \right\} = \frac{V_2}{V_1}
1482 \end{equation}
1483
1484 so that, with a constant-pressure gas thermometer using a gas for
1485 which the Joule-Thomson coefficient is zero, the Kelvin temperature is
1486 proportional to the measured volume.
1487
1488 Now consider another empirical fact, [[http://en.wikipedia.org/wiki/Boyle%27s_law][Boyle's law]]. For gases
1489 sufficiently far from their condensation points---which is also
1490 the condition under which (1-32) is satisfied---Boyle found that
1491 the product $PV$ is a constant at any fixed temperature. This
1492 product is, of course proportional to the number of moles $n$
1493 present, and so Boyle's equation of state takes the form
1494
1495 \begin{equation}PV = n \cdot f(t)\end{equation}
1496
1497 where f(t) is a function that depends on the particular empirical
1498 temperature scale used. But from (1-33) we must then have
1499 $f(t) = RT$, where $R$ is a constant, the universal gas constant whose
1500 numerical value (1.986 calories per mole per degree K) , depends
1501 on the size of the units in which we choose to measure the Kelvin
1502 temperature $T$. In terms of the Kelvin temperature, the ideal gas
1503 equation of state is therefore simply
1504
1505 \begin{equation}
1506 PV = nRT
1507 \end{equation}
1508
1509
1510 The relations (1-32) and (1-44) were found empirically, but
1511 with the development of thermodynamics one could show that they
1512 are not logically independent. In fact, all the material needed
1513 for this demonstration is now at hand, and we leave it as an
1514 exercise for the reader to prove that Joule‘s relation (1-32) is
1515 a logical consequence of Boyle's equation of state (1-44) and the
1516 first law.
1517
1518
1519 Historically, the advantages of the gas thermometer were
1520 discovered empirically before the Kelvin temperature scale was
1521 defined; and the temperature scale \theta defined by
1522
1523 \begin{equation}
1524 \theta = \lim_{P\rightarrow 0}\left(\frac{PV}{nR}\right)
1525 \end{equation}
1526
1527 was found to be convenient, easily reproducible, and independent
1528 of the properties of any particular gas. It was called the
1529 /absolute/ temperature scale; and from the foregoing it is clear
1530 that with the same choice of the numerical constant $R$, the
1531 absolute and Kelvin scales are identical.
1532
1533
1534 For many years the unit of our temperature scale was the
1535 Centigrade degree, so defined that the difference $T_b - T_f$ of
1536 boiling and freezing points of water was exactly 100 degrees.
1537 However, improvements in experimental techniques have made another
1538 method more reproducible; and the degree was redefined by the
1539 Tenth General Conference of Weights and Measures in 1954, by
1540 the condition that the triple point of water is at 273.l6^\circ K,
1541 this number being exact by definition. The freezing point, 0^\circ C,
1542 is then 273.15^\circ K. This new degree is called the Celsius degree.
1543 For further details, see the U.S. National Bureau of Standards
1544 Technical News Bulletin, October l963.
1545
1546
1547 The appearance of such a strange and arbitrary-looking
1548 number as 273.16 in the /definition/ of a unit is the result of
1549 the historical development, and is the means by which much
1550 greater confusion is avoided. Whenever improved techniques make
1551 possible a new and more precise (i.e., more reproducible)
1552 definition of a physical unit, its numerical value is of course chosen
1553 so as to be well inside the limits of error with which the old
1554 unit could be defined. Thus the old Centigrade and new Celsius
1555 scales are the same, within the accuracy with which the
1556 Centigrade scale could be realized; so the same notation, ^\circ C, is used
1557 for both . Only in this way can old measurements retain their
1558 value and accuracy, without need of corrections every time a
1559 unit is redefined.
1560
1561 #capitalize Joules?
1562 Exactly the same thing has happened in the definition of
1563 the calorie; for a century, beginning with the work of Joule,
1564 more and more precise experiments were performed to determine
1565 the mechanical equivalent of heat more and more accurately . But
1566 eventually mechanical and electrical measurements of energy be
1567 came far more reproducible than calorimetric measurements; so
1568 recently the calorie was redefined to be 4.1840 Joules, this
1569 number now being exact by definition. Further details are given
1570 in the aforementioned Bureau of Standards Bulletin.
1571
1572
1573 The derivations of this section have shown that, for any
1574 particular substance, there is (except for choice of units) only
1575 one temperature scale $T$ with the property that $dQ = TdS$ where
1576 $dS$ is the exact differential of some state function $S$. But this
1577 in itself provides no reason to suppose that the /same/ Kelvin
1578 scale will result for all substances; i.e., if we determine a
1579 \ldquo{}helium Kelvin temperature\rdquo{} and a
1580 \ldquo{}carbon dioxide Kelvin temperature\rdquo{} by the measurements
1581 indicated in (1-38), and choose the units so that they agree numerically at one point, will they then
1582 agree at other points? Thus far we have given no reason to
1583 expect that the Kelvin scale is /universal/, other than the empirical
1584 fact that the limit (1-45) is found to be the same for all gases.
1585 In section 2.0 we will see that this universality is a conse
1586 quence of the second law of thermodynamics (i.e., if we ever
1587 find two substances for which the Kelvin scale as defined above
1588 is different, then we can take advantage of this to make a
1589 perpetual motion machine of the second kind).
1590
1591
1592 Usually, the second law is introduced before discussing
1593 entropy or the Kelvin temperature scale. We have chosen this
1594 unusual order so as to demonstrate that the concepts of entropy
1595 and Kelvin temperature are logically independent of the second
1596 law; they can be defined theoretically, and the experimental
1597 procedures for their measurement can be developed, without any
1598 appeal to the second law. From the standpoint of logic, there
1599 fore, the second law serves /only/ to establish that the Kelvin
1600 temperature scale is the same for all substances.
1601
1602
1603 ** COMMENT Entropy of an Ideal Boltzmann Gas
1604
1605 At the present stage we are far from understanding the physical
1606 meaning of the function $S$ defined by (1-19); but we can investigate its mathematical
1607 form and numerical values. Let us do this for a system con
1608 sisting cf n moles of a substance which obeys the ideal gas
1609 equation of state
1610 and for which the heat capacity at constant volume CV is a
1611 constant. The difference in entropy between any two states (1)
1612 and (2) is from (1-19),
1613
1614
1615 where we integrate over any reversible path connecting the two
1616 states. From the manner in which S was defined, this integral
1617 must be the same whatever path we choose. Consider, then, a
1618 path consisting of a reversible expansion at constant tempera
1619 ture to a state 3 which has the initial temperature T, and the
1620 .L ' "'1 final volume V2; followed by heating at constant volume to the final temperature T2. Then (1-47) becomes
1621 3 2 I If r85 - on - db — = d — -4 S2 51 J V [aT]v M (1 8)
1622 1 3
1623 To evaluate the integral over (1 +3) , note that since
1624 dU = T :15 — P dV, the Helmholtz free energy function F E U — TS
1625 has the property dF = --S - P 61V; and of course is an exact
1626 differential since F is a definite state function. The condition
1627 that dF be exact is, analogous to (1-22),
1628 which is one of the Maxwell relations, discussed further in
1629 where CV is the molar heat capacity at constant volume. Collec
1630 ting these results, we have
1631 3
1632 l 3
1633 1 nR log(V2/V1) + nCV log(T2/Tl) (1-52)
1634 since CV was assumed independent of T. Thus the entropy function
1635 must have the form
1636 S(n,V,T) = nR log V + n CV log T + (const.) (l~53)
1637
1638
1639 From the derivation, the additive constant must be independent
1640 of V and T; but it can still depend on n. We indicate this by
1641 writing
1642 where f (n) is a function not determined by the definition (1-47).
1643 The form of f (n) is , however, restricted by the condition that
1644 the entropy be an extensive quantity; i .e . , two identical systems
1645 placed together should have twice the entropy of a single system;
1646 Substituting (l—-54) into (1-55), we find that f(n) must satisfy
1647 To solve this, one can differentiate with respect to q and set
1648 q = 1; we then obtain the differential equation
1649 n f ' (n) — f (n) + Rn = 0 (1-57)
1650 which is readily solved; alternatively, just set n = 1 in (1-56)
1651 and replace q by n . By either procedure we find
1652 f (n) = n f (1) — Rn log n . (1-58)
1653 As a check, it is easily verified that this is the solution of
1654 where A E f (l) is still an arbitrary constant, not determined
1655 by the definition (l—l9) , or by the condition (l-55) that S be
1656 extensive. However, A is not without physical meaning; we will
1657 see in the next Section that the vapor pressure of this sub
1658 stance (and more generally, its chemical potential) depends on
1659 A. Later, it will appear that the numerical value of A involves
1660 Planck's constant, and its theoretical determination therefore
1661 requires quantum statistics .
1662 We conclude from this that, in any region where experi
1663 mentally CV const. , and the ideal gas equation of state is
1664
1665
1666 obeyed, the entropy must have the form (1-59) . The fact that
1667 classical statistical mechanics does not lead to this result,
1668 the term nR log (l/n) being missing (Gibbs paradox) , was his
1669 torically one of the earliest clues indicating the need for the
1670 quantum theory.
1671 In the case of a liquid, the volume does not change appre
1672 ciably on heating, and so d5 = n CV dT/T, and if CV is indepen
1673 dent of temperature, we would have in place of (1-59) ,
1674 where Ag is an integration constant, which also has physical
1675 meaning in connection with conditions of equilibrium between
1676 two different phases.
1677 1.1.0 The Second Law: Definition. Probably no proposition in
1678 physics has been the subject of more deep and sus tained confusion
1679 than the second law of thermodynamics . It is not in the province
1680 of macroscopic thermodynamics to explain the underlying reason
1681 for the second law; but at this stage we should at least be able
1682 to state this law in clear and experimentally meaningful terms.
1683 However, examination of some current textbooks reveals that,
1684 after more than a century, different authors still disagree as
1685 to the proper statement of the second law, its physical meaning,
1686 and its exact range of validity.
1687 Later on in this book it will be one of our major objectives
1688 to show, from several different viewpoints , how much clearer and
1689 simpler these problems now appear in the light of recent develop
1690 ments in statistical mechanics . For the present, however, our
1691 aim is only to prepare the way for this by pointing out exactly
1692 what it is that is to be proved later. As a start on this at
1693 tempt, we note that the second law conveys a certain piece of
1694 informations about the direction in which processes take place.
1695 In application it enables us to predict such things as the final
1696 equilibrium state of a system, in situations where the first law
1697 alone is insufficient to do this.
1698 A concrete example will be helpful. We have a vessel
1699 equipped with a piston, containing N moles of carbon dioxide.
1700
1701
1702 The system is initially at thermal equilibrium at temperature To, volume V0 and pressure PO; and under these conditions it contains
1703 n moles of CO2 in the vapor phase and moles in the liquid
1704 phase . The system is now thermally insulated from its surround
1705 ings, and the piston is moved rapidly (i.e. , so that n does not
1706 change appreciably during the motion) so that the system has a
1707 new volume Vf; and immediately after the motion, a new pressure
1708 PI . The piston is now held fixed in its new position , and the
1709 system allowed to come once more to equilibrium. During this
1710 process, will the CO2 tend to evaporate further, or condense further? What will be the final equilibrium temperature Teq, the final pressure PeCE , and final value of n eq?
1711 It is clear that the firs t law alone is incapable of answering
1712 these questions; for if the only requirement is conservation of
1713 energy, then the CO2 might condense , giving up i ts heat of vapor
1714 ization and raising the temperature of the system; or it might
1715 evaporate further, lowering the temperature. Indeed, all values
1716 of neq in O i neq i N would be possible without any violation of the first law. In practice, however, this process will be found
1717 to go in only one direction and the sys term will reach a definite
1718 final equilibrium state with a temperature, pressure, and vapor
1719 density predictable from the second law.
1720 Now there are dozens of possible verbal statements of the
1721 second law; and from one standpoint, any statement which conveys
1722 the same information has equal right to be called "the second
1723 law." However, not all of them are equally direct statements of
1724 experimental fact, or equally convenient for applications, or
1725 equally general; and it is on these grounds that we ought to
1726 choose among them .
1727 Some of the mos t popular statements of the s econd law be
1728 long to the class of the well-—known "impossibility" assertions ;
1729 i.e. , it is impossible to transfer heat from a lower to a higher
1730 temperature without leaving compensating changes in the rest of
1731 the universe , it is imposs ible to convert heat into useful work
1732 without leaving compensating changes, it is impossible to make
1733 a perpetual motion machine of the second kind, etc.
1734
1735
1736 Suoh formulations have one clear logical merit; they are
1737 stated in such a way that, if the assertion should be false, a
1738 single experiment would suffice to demonstrate that fact conclu
1739 sively. It is good to have our principles stated in such a
1740 clear, unequivocal way.
1741 However, impossibility statements also have some disadvan
1742 tages . In the first place, their_ are not, and their very
1743 nature cannot be, statements of eiperimental fact. Indeed, we
1744 can put it more strongly; we have no record of anyone having
1745 seriously tried to do any of the various things which have been
1746 asserted to be impossible, except for one case which actually
1747 succeeded‘. In the experimental realization of negative spin
1748 temperatures , one can transfer heat from a lower to a higher
1749 temperature without external changes; and so one of the common
1750 impossibility statements is now known to be false [for a clear
1751 discussion of this, see the article of N. F . Ramsey (1956) ;
1752 experimental details of calorimetry with negative temperature
1753 spin systems are given by Abragam and Proctor (1958) ] .
1754 Finally, impossibility statements are of very little use in
1755 applications of thermodynamics; the assertion that a certain kind
1756 of machine cannot be built, or that a -certain laboratory feat
1757 cannot be performed, does not tell me very directly whether my
1758 carbon dioxide will condense or evaporate. For applications,
1759 such assertions must first be converted into a more explicit
1760 mathematical form.
1761 For these reasons, it appears that a different kind of
1762 statement of the second law will be, not necessarily more
1763 "correct,” but more useful in practice. Now both Clausius (3.875)
1764 and Planck (1897) have laid great stress on their conclusion
1765 that the most general statement, and also the most immediately
1766 useful in applications, is simply the existence of a state
1767 function, called the entropy, which tends to increase. More
1768 precisely: in an adiabatic change of state, the entropy of
1769 a system may increase or may remain constant, but does not
1770 decrease. In a process involving heat flow to or from the
1771 system, the total entropy of all bodies involved may increase
1772
1773
1774 or may remain constant; but does not decrease; let us call this
1775 the “weak form" of the second law.
1776 The weak form of the second law is capable of answering the
1777 first question posed above; thus the carbon dioxide will evapo
1778 rate further if , and only if , this leads to an increase in the
1779 total entropy of the system . This alone , however , is not enough
1780 to answer the second question; to predict the exact final equili
1781 brium state, we need one more fact.
1782 The strong form of the second law is obtained by adding the
1783 further assertion that the entropy not only “tends" to increase;
1784 in fact it will increase, to the maximum value permitted E2 the
1785 constraints imposed.* In the case of the carbon dioxide, these
1786 constraints are: fixed total energy (first law) , fixed total
1787 amount of carbon dioxide , and fixed position of the piston . The
1788 final equilibrium state is the one which has the maximum entropy
1789 compatible with these constraints , and it can be predicted quan
1790 titatively from the strong form of the second law if we know,
1791 from experiment or theory, the thermodynamic properties of carbon
1792 dioxide (i .e . , heat capacity , equation of state , heat of vapor
1793 ization) .
1794 To illus trate this , we set up the problem in a crude ap
1795 proximation which supposes that (l) in the range of conditions
1796 of interest, the molar heat capacity CV of the vapor, and C2 of
1797 the liquid, and the molar heat of vaporization L, are all con
1798 stants, and the heat capacities of cylinder and piston are neg
1799 ligible; (2) the liquid volume is always a small fraction of the
1800 total V, so that changes in vapor volume may be neglected; (3) the
1801 vapor obeys the ideal gas equation of state PV = nRT. The in
1802 ternal energy functions of liquid and vapor then have the form
1803 UPb = + A} (1-61)
1804 T T U = n‘ C '1‘ A + L] (1-62)
1805 v , v
1806 where A is a constant which plays no role in the problem. The
1807 appearance of L in (1-62) recognizes that the zero from which we
1808 *Note , however , that the second law has nothing to say about how rapidly this approach to equilibrium takes place.
1809
1810
1811 measure energy of the vapor is higher than that of the liquid by
1812 the energy L necessary to form the vapor. On evaporation of dn
1813 moles of liquid, the total energy increment is (ill = + dUV= O,
1814 or
1815 [n CV [(CV — CQ)T + = O (l—63)
1816 which is the constraint imposed by the first law. As we found
1817 previously (l~59) , (1-60) the entropies of vapor and liquid are
1818 given by
1819 S = n [C 1n T + R ln (V/n) + A ] (1-64)
1820 v v v
1821 where AV, ASL are the constants of integration discussed in the
1822 Si
1823 last Section.
1824 We leave it as an exercise for the reader to complete the
1825 derivation from this point , and show that the total entropy
1826 S = 82 + SV is maximized subject to the constraint (1-6 3) , when
1827 R
1828 the values 11 , T are related by
1829 eq eq
1830 Equation (1-66) is recognized as an approximate form of the Vapor
1831 pressure formula .
1832 We note that AQ, AV, which appeared first as integration
1833 constants for the entropy with no parti cular physical meaning ,
1834 now play a role in determining the vapor pressure.
1835 l.ll The Second Law: Discussion. We have emphasized the dis
1836 tinction between the weak and strong forms of the second law
1837 because (with the exception of Boltzmann ' s original unsuccessful
1838 argument based on the H—theorem) , most attempts to deduce the
1839 second law from statis tical mechanics have considered only the
1840 weak form; whereas it is evidently the strong form that leads
1841 to definite quantitative predictions, and is therefore needed
908 1842
909 1843
910 1844
911 * COMMENT Appendix 1845 * COMMENT Appendix
912 1846