0810.1019/E1.tex
1: 
2: 
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \chapter{Phenomenological thermodynamics}\label{c.ctherm}
5: 
6: Part II discusses statistical mechanics from an algebraic perspective,
7: concentrating on thermal equilibrium but discussing basic things in 
8: a more general framework.
9: A treatment of equilibrium statistical mechanics
10: and the kinematic part of nonequilibrium statistical mechanics
11: is given which derives from a single basic assumption (Definition 
12: \ref{3.1.}) the full structure of phenomenological thermodynamics and 
13: of statistical mechanics, except for the third law 
14: which requires an additional quantization assumption.
15: 
16: \bigskip
17: This chapter gives a concise description of standard phenomenological 
18: equilibrium thermodynamics for single-phase systems in the absence 
19: of chemical reactions and electromagnetic fields. 
20: From the formulas provided, it is an easy step to go to various 
21: examples and applications discussed in standard textbooks such as
22: \sca{Callen} \cite{Cal} or \sca{Reichl} \cite{Rei}.
23: A full discussion of global equilibrium would also involve the
24: equilibrium treatment of multiple phases and chemical reactions. 
25: Since their discussion offers no new aspects compared with 
26: traditional textbook treatments, they are not treated here.
27: 
28: Our phenomenological approach is similar to that of \sca{Callen} 
29: \cite{Cal}, who introduces the basic concepts by means of a few 
30: postulates from which everything else follows. 
31: The present setting is a modified version designed to
32: match the more fundamental approach based on statistical 
33: mechanics. By specifying the kinematical properties of states 
34: outside equilibrium, his informal thermodynamic 
35: stability arguments (which depends on a dynamical assumption close to 
36: equilibrium) can be replaced by rigorous mathematical arguments.
37: 
38: 
39: 
40: 
41: 
42: \at{adapt Section 1.5 to match the contents}
43: 
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: \section{Standard thermodynamical systems}\label{s.phen}
46: 
47: 
48: We discuss here the special but very important case of thermodynamic 
49: systems describing the single-phase global equilibrium of matter 
50: composed of one or several kinds of substances in the absence of 
51: chemical reactions and electromagnetic fields. We call such systems 
52: \bfi{standard thermodynamic systems}; they are ubiquitous in 
53: applications. 
54: In particular, a standard system is considered to be uncharged, 
55: homogeneous, and isotropic, so that each finite 
56: region looks like any other and is very large in microscopic units.
57: 
58: The substances of fixed chemical composition are labeled by an index 
59: $j\in J$. A standard thermodynamic system is completely characterized 
60: by\footnote{
61: In the terminology, we mainly follow the IUPAC convention 
62: (\sca{Alberty} \cite[Section 7]{Alb}), except that we use the letter
63: $H$ to denote the Hamilton energy, as customary in quantum mechanics. 
64: In equilibrium, $H$ equals the internal energy $U$. The Hamilton energy 
65: should not be confused with the enthalpy which is usually denoted by 
66: $H$ but here is given in equilibrium by $H+PV$.
67: For a history of thermodynamics notation, see 
68: \sca{Battino} et al. \cite{BatSW}.
69: } % end footnote
70: the \bfi{mole number} $N_j$ of each substance $j$, the corresponding 
71: \bfi{chemical potential} $\mu_j$ of substance $j$, the \bfi{volume} $V$,
72: the \bfi{pressure} $P$, the \bfi{temperature} $T$, the \bfi{entropy} 
73: $S$, and the \bfi{Hamilton energy} $H$. These variables, the 
74: \bfi{extensive} variables $N_j,V,S,H$ and the \bfi{intensive} variables 
75: $\mu_j,P,T$, are jointly called the \bfi{basic thermodynamic variables}.
76: We group the $N_j$ and the $\mu_j$ into vectors $N$ and $\mu$ indexed 
77: by $J$ and write $\mu\cdot N = \sum_{j\in J} \mu_jN_j$. 
78: In the special case of a \bfi{pure substance}, there is just a single 
79: kind of substance; then we drop the indices and have 
80: $\mu\cdot N = \mu N$. In this section, all numbers are real.
81: 
82: The mathematics of thermodynamics makes essential use of the concept 
83: of convexity. A set $X\subseteq \Rz^n$ is called \bfi{convex} if 
84: $tx+(1-t)y\in X$  for all $x,y\in X$ and all $t\in[0,1]$.
85: A real-valued function $\phi$ is called \bfi{convex} in the convex set
86: $X\subseteq \Rz^n$ if $\phi$ is defined on $X$ and, for all $x,y\in X$,
87: \[
88: \phi(tx+(1-t)y) \le t\phi(x)+(1-t)\phi(y) \for 0\le t\le 1.
89: \]
90: Clearly, $\phi$ is convex iff for all $x,y\in X$, the function 
91: $\mu:[0,1]\to \Rz$ defined by
92: \[
93: \mu(t):=\phi(x+t(y-x))
94: \]
95: is convex. It is well-known that, for twice continuously 
96: differentiable $\phi$, this is the case iff the second derivative 
97: $\mu''(t)$ is nonnegative for $0\le t\le 1$. 
98: Note that by a theorem of Aleksandrov (see \sca{Aleksandrov} \cite{Ale},
99: \sca{Alberti \& Ambrosio} \cite{AlbA}, \sca{Rockafellar} \cite{Roc.ca}),
100: convex functions are almost everywhere twice continuously 
101: differentiable: For almost every $x\in X$, there exist a 
102: unique vector $\partial\phi(x)\in \Rz^n$, the \bfi{gradient} of $\phi$ 
103: at $x$, and a unique symmetric, positive semidefinite matrix 
104: $\partial^2\phi(x)\in \Rz^{n\times n}$, the \bfi{Hessian} of $\phi$ 
105: at $x$, such that 
106: \[
107: \phi(x+h)=\phi(x)+h^T\partial\phi(x)
108: +\half h^T\partial^2\phi(x)h + o(\|h\|^2)
109: \]
110: for sufficiently small $h \in \Rz^n$.
111: A function $\phi$ is called \bfi{concave} if $-\phi$ is convex. 
112: Thus, for a twice continuously differentiable function $\phi$ of a 
113: single variable $\tau$, $\phi$ is concave iff $\mu''(\tau)\le 0$ for 
114: $0\le \tau\le 1$. 
115: 
116: \begin{prop}\label{e.convex}
117: If $\phi$ is convex in the convex set $X$ then the function $\psi$ 
118: defined by
119: \[
120: \psi(s,x):=s\phi(x/s)
121: \]
122: is convex in the set $\{(s,x)\in \Rz x X \mid s> 0\}$ 
123: and concave in the set $\{(s,x)\in \Rz x X \mid s< 0\}$. 
124: \end{prop}
125: 
126: \bepf
127: It suffices to show that $\mu(t):=\psi(s+tk,x+th)$ is convex (concave)
128: for all $s,x,h,k$ such that $s+tk>0$ (resp. $<0$).
129: Let $z(t):=(x+th)/(s+tk)$ and $c:=sh-kx$. Then
130: \[
131: z'(t)=\frac{c}{(s+tk)^2},~~~\mu(t)=(s+tk)\phi(z(t)),
132: \]
133: hence
134: \[
135: \mu'(t)=k\phi(z(t))+\phi'(z(t))\frac{c}{s+tk},
136: \]
137: \[
138: \mu''(t)=k\phi'(z(t))\frac{c}{(s+tk)^2}
139: +\frac{c^T}{(s+tk)^2}\phi''(z(t))\frac{c}{s+tk}
140: +\phi'(z(t))\frac{-ck}{(s+tk)^2}
141: =\frac{c^T\phi''(z(t))c}{(s+tk)^3},
142: \]
143: which has the required sign.
144: \epf
145: 
146: 
147: Equilibrium thermodynamics is about characterizing so-called
148: equilibrium states in terms of intensive and extensive variables 
149: and their relations, and comparing them with similar nonequilibrium 
150: states. In a nonequilibrium state, only extensive variables have a 
151: well-defined meaning; but these are not sufficient to characterize 
152: system behavior completely.
153: 
154: All valid statements in the equilibrium thermodynamics of standard
155: systems can be deduced from the following definition.
156: 
157: \begin{dfn}\bfi{(Phenomenological thermodynamics)}
158: \label{d.phen}\\
159: (i) Temperature $T$, pressure $P$, and volume $V$ are positive,
160: mole numbers $N_j$ are nonnegative. 
161: The extensive variables $H,S,V,N$ are additive under the 
162: composition of disjoint subsystems.
163: 
164: (ii) There is a convex \bfi{system function} $\Delta$ of the intensive 
165: variables $T,P,\mu$ which is monotone increasing in $T$ and monotone 
166: decreasing in $P$, such that the intensive variables are related by 
167: the \bfi{equation of state}
168: \lbeq{e.def}
169: \Delta(T,P,\mu)=0.
170: \eeq
171: The set of $(T,P,\mu)$ satisfying $T>0$, $P>0$ and the equation of 
172: state is called the \bfi{state space}. 
173: 
174: (iii) The Hamilton energy $H$ satisfies 
175: the \bfi{Euler inequality}
176: \lbeq{e.Hi}
177: H\ge T S - P V + \mu \cdot N
178: \eeq
179: for all $(T,P,\mu)$ in the state space.
180: \bfi{Equilibrium states} have well-defined intensive and extensive 
181: variables satisfying equality in \gzit{e.Hi}.
182: A system is in \bfi{equilibrium} if it is completely described by an 
183: equilibrium state.
184: \end{dfn}
185: 
186: This is the complete list of assumptions defining phenomenological 
187: equilibrium thermodynamics; the system function $\Delta$ can be 
188: determined either by fitting to experimental data, or by calculation 
189: from a more fundamental description, cf. Theorem \ref{t.eos}.
190: All other properties follow from the system function.
191: Thus, all equilibrium properties of a material are characterized by 
192: the system function $\Delta$.
193: 
194: Surfaces of nondifferentiability of the system function
195: correspond to so-called \bfi{phase transitions}. 
196: The equation of state shows that, apart from possible phase 
197: transitions, the state space has the 
198: structure of an $(s-1)$-dimensional manifold in $\Rz ^{s}$, where 
199: $s$ is the number of intensive variables; in case of a standard 
200: system, the dimension is therefore one higher than the number of kinds 
201: of substances.
202: 
203: Thermodynamic systems with multiple phases are only piecewise 
204: homogeneous; each phase separately may be described 
205: as a standard thermodynamic system, but discussing the equilibrium 
206: at interfaces needs some additional effort, described in all textbooks 
207: on thermodynamics. Therefore, we consider only regions 
208: of state space where the system function $\Delta$ is twice continuously 
209: differentiable.
210: 
211: 
212: \bigskip
213: Each equilibrium instance of the 
214: material is characterized by a particular state 
215: $(T,P,N)$, from which all equilibrium properties can be computed:
216: 
217: \begin{thm}~
218: 
219: (i) In any equilibrium state, the extensive variables are given by 
220: \lbeq{e.Sp}
221: S=\Omega\frac{\partial \Delta}{\partial T}(T,P,\mu),~~~
222: V=-\Omega\frac{\partial \Delta}{\partial P}(T,P,\mu),~~~
223: N=\Omega\frac{\partial \Delta}{\partial \mu}(T,P,\mu),
224: \eeq
225: and the \bfi{Euler equation}
226: \lbeq{e.Hp}
227: H=T S - P V + \mu \cdot N.
228: \eeq
229: Here $\Omega$ is a positive number called the \bfi{system size}. 
230: 
231: (ii) In equilibrium, we have the \bfi{Maxwell reciprocity relations}
232: \lbeq{7-10p}
233: -\frac {\partial V} {\partial T}
234: =\frac {\partial S} {\partial P},~~
235: \frac {\partial N_j} {\partial T}
236: =\frac {\partial S} {\partial\mu_j},~~
237: \frac {\partial N_j} {\partial P}
238: =-\frac {\partial V} {\partial\mu_j},~~
239: \frac {\partial N_j} {\partial\mu_k}
240: =\frac {\partial N_k} {\partial\mu_j}
241: \eeq
242: and the \bfi{stability conditions}
243: \lbeq{7-13p}
244: \frac {\partial S} {\partial T}\ge 0,~~ 
245: \frac {\partial V} {\partial P}\le 0,~~ 
246: \frac {\partial N_j} {\partial\mu_j}\ge 0.
247: \eeq
248: \end{thm}
249: 
250: \bepf
251: At fixed $S,V,N$, inequality \gzit{e.Hi} holds in equilibrium with 
252: equality. Therefore the triple $(T,P,\mu)$ is a maximizer
253: of $TS-PV+\mu\cdot N$ under the constraints $\Delta(T,P,\mu)=0$, $T>0$, 
254: $P>0$. A necessary condition for a maximizer is the stationarity of 
255: the Lagrangian
256: \[
257: L(T,P,\mu)=TS-PV+\mu\cdot N -\Omega\Delta(T,P,\mu)
258: \]
259: for some Lagrange multiplier $\Omega$. Setting the partial derivatives 
260: to zero gives \gzit{e.Sp}, and since the maximum is attained in 
261: equilibrium, the Euler equation \gzit{e.Hp} follows. The system size
262: $\Omega$ is positive since $V>0$ and $\Delta$ is decreasing in $P$.
263: Since the Hessian matrix of $\Delta$, 
264: \[
265: \Sigma= 
266: \left(\begin{array}{rrr}
267: \D\frac{\partial^2\Delta}{\partial T^2}& 
268: \D\frac{\partial^2\Delta}{\partial P\partial T}& 
269: \D\frac{\partial^2\Delta}{\partial \mu\partial T} \\
270: % ~\\
271: \D\frac{\partial^2\Delta}{\partial T\partial P}& 
272: \D\frac{\partial^2\Delta}{\partial P^2}& 
273: \D\frac{\partial^2\Delta}{\partial \mu\partial P} \\
274: % ~\\
275: \D\frac{\partial^2\Delta}{\partial T \partial \mu}& 
276: \D\frac{\partial^2\Delta}{\partial P \partial \mu}& 
277: \D\frac{\partial^2\Delta}{\partial\mu^2}
278: \end{array}\right)
279: = \Omega^{-1}
280: \left(\begin{array}{rrr}
281: \D\frac{\partial S}{\partial T}& 
282: \D\frac{\partial S}{\partial P}& 
283: \D\frac{\partial S}{\partial \mu} \\
284: % ~\\
285: \D-\frac{\partial V}{\partial T}& 
286: \D-\frac{\partial V}{\partial P}& 
287: \D-\frac{\partial V}{\partial \mu} \\
288: % ~\\
289: \D\frac{\partial N}{\partial T}& 
290: \D\frac{\partial N}{\partial P}& 
291: \D\frac{\partial N}{\partial\mu}
292: \end{array}\right),
293: \]
294: is symmetric, the Maxwell reciprocity relations follow. Since 
295: $\Delta$ is convex, $\Sigma$ is positive semidefinite; hence the 
296: diagonal elements of $\Sigma$ are nonnegative, giving the stability 
297: conditions.
298: \epf
299: 
300: Note that there are further stability conditions since the determinants
301: of all principal submatrices of $\Sigma$ must be nonnegative. 
302: In addition, $N_j\ge 0$ implies that $\Delta$ is monotone increasing 
303: in each $\mu_j$.
304: 
305: \begin{expl}\label{ex.ideal}
306: The equilibrium behavior of electrically neutral gases at sufficiently 
307: low pressure can be modelled as ideal gases.
308: An \bfi{ideal gas} is defined by a system function of the form
309: \lbeq{e.ideal}
310: \Delta(T,P,\mu)=\sum_{j\in J}\pi_j(T) e^{\mu_j/ R T}-P,
311: \eeq
312: where the $\pi_j(T)$ are positive functions of the temperature,
313: \lbeq{e.R}
314: R\approx 8.31447\, JK^{-1}\mbox{mol}^{-1}
315: \eeq
316: is the \bfi{universal gas constant}\footnote{
317: For the internationally recommended values of this and other constants, 
318: their accuracy, determination, and history, 
319: see \sca{CODATA} \cite{CODATA}.
320: }, % end footnote
321: and we use the bracketing convention $\mu_j/ R T =\mu_j/( R T)$.
322: Differentiation with respect to $P$ shows that $\Omega=V$ is the 
323: system size, and from \gzit{e.def}, \gzit{e.Sp}, and equality in 
324: \gzit{e.Hi}, we find that, in equilibrium,
325: \[
326: P=\sum_j \pi_j(T) e^{\mu_j/ R T},~~~
327: S=V\sum_j \Big(\frac{\partial}{\partial T}\pi_j(T) 
328: -\frac{\mu_j\pi_j(T)}{ R T^2}\Big) e^{\mu_j/ R T},
329: \]
330: \[
331: N_j =\frac{V\pi_j(T)}{ R T} e^{\mu_j/ R T},~~~
332: H=V\sum_j \Big(T\frac{\partial}{\partial T}\pi_j(T) 
333: -\pi_j(T)\Big) e^{\mu_j/ R T}.
334: \]
335: Expressed in terms of $T,V,N$, we have
336: \[
337: PV= R T \sum_j N_j,~~~\mu_j= R T\log\frac{ R T N_j}{V\pi_j(T)},
338: \]
339: \[
340: H=\sum_j h_j(T)N_j,~~~
341: h_j(T)= R T\Big(T\frac{\partial}{\partial T}\log\pi_j(T) -1\Big),
342: \]
343: from which $S$ can be computed by means of the Euler equation 
344: \gzit{e.Hp}. 
345: In particular, for one \bfi{mole} of a single substance, defined by 
346: $N=1$, we get the \bfi{ideal gas law} 
347: \lbeq{e.ilaw}
348: PV=RT
349: \eeq
350: discovered by \sca{Clapeyron} \cite{Clap}; cf. \sca{Jensen} \cite{Jen}.
351: 
352: 
353: In general, the difference $h_j(T)-h_j(T')$ can be found 
354: experimentally by measuring the energy needed for raising or lowering 
355: the temperature of pure substance $j$ from $T'$ to $T$ while keeping 
356: the $N_j$ constant. In terms of infinitesimal increments, the 
357: \bfi{heat capacities} 
358: \[
359: C_j(T)=dh_j(T)/dT,
360: \]
361: we have
362: \[
363: h_j(T)=h_j(T')+\int_{T'}^T dT\, C_j(T).
364: \]
365: From the definition of $h_j(T)$, we find that
366: \[
367: \pi_j(T)=\pi_j(T')\exp \int_{T'}^T \frac{dT}{T}
368: \Big(1+\frac{h_j(T)}{ R T}\Big).
369: \]
370: Thus there are two undetermined integration constants for each kind of
371: substance. These cannot be determined experimentally as long
372: as we are in the range of validity of the ideal gas approximation.
373: Indeed, if we pick arbitrary constants $\alpha_j$ and $\gamma_j$ and 
374: replace $\pi_j(T),\mu_j,H$, and $S$ by 
375: \[
376: \pi_j'(T):=e^{\alpha_j-\gamma_j/ R T}\pi_j(T),~~~
377: \mu_j'=\mu_j+\gamma_j- R T \alpha_j,
378: \]
379: \[
380: H'=H+\sum_j \alpha_jN_j,~~~S'=S+ R \sum_j \gamma_jN_j,
381: \]
382: all relations remain unchanged. Thus, the Hamilton energy and the 
383: entropy of an ideal gas are only determined up to an arbitrary linear 
384: combination of the mole numbers. This is an instance of the deeper 
385: problem to determine under which conditions thermodynamic variables 
386: are controllable; cf. the discussion in the context of Example 
387: \ref{ex.gibbs} below. 
388: 
389: The gauge freedom 
390: (present only in the ideal gas) can be fixed by choosing a particular 
391: \bfi{standard temperature} $T_0$ and setting arbitrarily $h_j(T_0)=0$, 
392: $\mu_j(T_0)=0$. 
393: Alternatively, at sufficiently large temperature $T$, heat capacities 
394: are usually nearly constant, and making use of the gauge freedom, 
395: we may simply assume that 
396: \[
397: h_j(T)=h_{j0} T,~~~\pi_j(T)=\pi_{j0} T \mbox{~~~for large~} T.
398: \]
399: Note that this gauge freedom is present only for ideal gases. 
400: \end{expl}
401: 
402: 
403: 
404: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
405: \section{The laws of thermodynamics}\label{s.cons}
406: 
407: The possibility of measuring temperature is sometimes called the
408: \bfi{zeroth law of thermodynamics} (\sca{Fowler \& Guggenheim}
409: \cite{FowG}). For the history of temperature, see 
410: \sca{Roller} \cite{Rol} and \sca{Truesdell} \cite{Tru}.
411: The ideal gas law \gzit{e.ilaw} is the basis for the 
412: construction of a \bfi{gas thermometer}: The amount of expansion of 
413: volume in a long, thin tube can easily be read off from a scale along 
414: the tube. 
415: We have $V=aL$, where $a$ is the cross section area and $L$ is the 
416: length of the filled part of the tube, hence $T=(aP/R)L$. Thus, at 
417: constant pressure, the temperature of the gas is proportional to $L$. 
418: 
419: We say that two thermodynamic systems are brought in good \bfi{thermal 
420: contact} if the joint system tends after a short time to an equilibrium 
421: state. To measure the temperature of a system,
422: one brings it in thermal contact with a thermometer and waits until 
423: equilibrium is established.
424: The system and the thermometer will then have the same temperature, 
425: which can be read off from the thermometer. If the system is much larger
426: than the thermometer, this temperature will be essentially the same 
427: as the temperature of the system before the measurement.
428: For a survey of the problems involved in defining and
429: measuring temperature outside equilibrium, see 
430: \sca{Casas-V\'asquez \& Jou} \cite{CasJ}.
431: 
432: \bigskip
433: To be able to formulate the first law of thermodynamics we need the
434: concept of a reversible change of states, i.e., changes
435: preserving the equilibrium condition. For use in later sections,
436: we define the concept in a slightly more general form,
437: writing $\alpha$ for $P$ and $\mu$ jointly.
438: 
439: \begin{dfn}
440: A \bfi{state variable} is an almost everywhere continuously 
441: differentiable function $\phi(T,\alpha)$ defined on the
442: state space (or a subset of it). 
443: Temporal changes in a state variable that 
444: occur when the boundary conditions are kept fixed are called 
445: \bfi{spontaneous changes}. 
446: A \bfi{reversible transformation} is a continuously differentiable 
447: mapping 
448: \[
449: \lambda \to (T(\lambda ),\alpha(\lambda ))
450: \]
451: from a real interval into the state space; thus
452: $\Delta(T(\lambda ),\alpha(\lambda ))=0$. The \bfi{differential}
453: \lbeq{3-12}
454: d\phi=\frac {\partial \phi} {\partial T}dT
455: +\frac {\partial \phi} {\partial \alpha} \cdot d\alpha, 
456: \eeq
457: obtained by multiplying the chain rule by $d\lambda$,
458: describes the change of a state variable $\phi$ under 
459: arbitrary (infinitesimal) reversible transformations. 
460: In formal mathematical terms, differentials are exact linear forms on 
461: the state space manifold; cf. Chapter \ref{c.manifolds}.
462: \end{dfn}
463: 
464: Reversible changes per se have nothing to do with changes in time. 
465: However, by sufficiently slow, quasistatic changes of the boundary 
466: conditions, reversible changes can often be realized approximately as 
467: temporal changes. The degree to which this is possible determines the 
468: efficiency of thermodynamic machines. The analysis of the efficiency
469: by means of the so-called \bfi{Carnot cycle} was the historical origin 
470: of thermodynamics.
471: 
472: The state space is often parameterized by 
473: different sets of state variables, as required by the application. 
474: If $T=T(\kappa,\lambda)$, $\alpha=\alpha(\kappa,\lambda)$ is such a
475: parameterization then the state variable $g(T,\alpha)$ can be written
476: as a function of $(\kappa,\lambda)$,
477: \lbeq{e.partial0}
478: g(\kappa,\lambda) = g(T(\kappa,\lambda),\alpha(\kappa,\lambda)).
479: \eeq
480: This notation, while mathematically ambiguous, is common in the
481: literature; the names of the argument decide which function is intended.
482: When writing partial derivatives without arguments, this leads to
483: serious ambiguities. These can be resolved by writing 
484: $\D\Big(\frac{\partial g}{\partial \lambda}\Big)_\kappa$ for the 
485: partial derivative of \gzit{e.partial0} with respect to $\lambda$;
486: it can be evaluated using \gzit{3-12}, giving the \bfi{chain rule}
487: \lbeq{e.partial}
488: \Big(\frac{\partial g}{\partial \lambda}\Big)_\kappa
489: =\frac{\partial g} {\partial T}
490: \Big(\frac{\partial T}{\partial \lambda}\Big)_\kappa
491: +\frac {\partial g} {\partial \alpha} \cdot
492: \Big(\frac{\partial \alpha}{\partial \lambda}\Big)_\kappa.
493: \eeq
494: Here the partial derivatives in the original 
495: parameterization by the intensive variables are written without 
496: parentheses.
497: 
498: Differentiating the equation of state \gzit{e.def}, using the chain 
499: rule \gzit{3-12}, and simplifying using \gzit{e.Sp} gives the 
500: \bfi{Gibbs-Duhem equation}
501: \lbeq{e.GDp}
502: 0=SdT- VdP+N\cdot d\mu
503: \eeq
504: for reversible changes. If we differentiate the 
505: Euler equation \gzit{e.Hp}, we obtain
506: \[
507: dH=TdS+SdT-PdV-VdP+\mu\cdot dN+N\cdot d\mu,
508: \]
509: and using \gzit{e.GDp}, this simplifies to the 
510: \bfi{first law of thermodynamics} 
511: \lbeq{3.1stp}
512: dH=TdS-Pd V +\mu \cdot dN.
513: \eeq
514: Historically, the first law of thermodynamics took on this form only 
515: gradually, through work by \sca{Mayer} \cite{May},
516: \sca{Joule}\cite{Jou}, \sca{Helmholtz}\cite{Hel}, and
517: \sca{Clausius} \cite{Cla}.
518: 
519: Considering global equilibrium from a fundamental point of view, the 
520: extensive variables are the variables that are conserved or at least 
521: change so slowly that they may be regarded as time independent on the 
522: time scale of interest. In the absence of chemical reactions, the 
523: mole numbers, the entropy, and the Hamilton energy are conserved; 
524: the volume is a system size variable which, in the fundamental view, 
525: must be taken as infinite (thermodynamic limit) to exclude the 
526: unavoidable interaction with the environment. However, real systems 
527: are always in contact with their environment,
528: and the conservation laws are approximate only. In thermodynamics, 
529: the description of the system boundary is generally reduced to the 
530: degrees of freedom observable at a given resolution.
531: 
532: The result of this reduced description (for derivations, see, e.g., 
533: \sca{Balian} \cite{Bal}, \sca{Grabert} \cite{Gra}, 
534: \sca{Rau \& M\"uller} \cite{RauM}) is a dynamical effect called 
535: \bfi{dissipation} (\sca{Thomson} \cite{Tho}). It is described by the 
536: \bfi{second law of thermodynamics}, which was discovered by
537: (\sca{Clausius} \cite{Cla2}.
538: The Euler inequality \gzit{e.Hi} together with the Euler equation 
539: \gzit{e.Hp} only express the nondynamical part of the second law since, 
540: in equilibrium thermodynamics, dynamical questions are ignored: 
541: Axiom (ii) says that if $S,V,N$ are conserved (thermal, mechanical and 
542: chemical isolation) then the \bfi{internal energy},
543: \lbeq{e.int}
544: U:=TS-PV+\mu\cdot N
545: \eeq
546: is minimal in equilibrium; 
547: \at{formulate more clearly; $T,P,\mu$ have no meaning outside
548:  equilibrium!}
549: if $T,V,N$ are conserved (mechanical and 
550: chemical isolation of a system at constant 
551: temperature $T$) then the \bfi{Helmholtz (free) energy},
552: \[
553: A:=U-TS=-PV+\mu\cdot N 
554: \]
555: is minimal in equilibrium; and if $T,P,N$ are conserved (chemical 
556: isolation of a system at constant temperature $T$ and pressure $P$) 
557: then the \bfi{Gibbs (free) energy},
558: \[
559: G:=A+PV=\mu\cdot N
560: \] 
561: is minimal in equilibrium.
562: 
563: The \bfi{third law of thermodynamics}, due to \sca{Nernst} \cite{Ner},
564: says that entropy is nonnegative. In view of \gzit{e.Sp}, 
565: this is equivalent to the monotonicity of $\Delta(T,P,\mu)$.
566: 
567: 
568: 
569: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
570: \section{Consequences of the first law}\label{s.c1}
571: 
572: The first law of thermodynamics describes the observable
573: energy balance in a reversible process. 
574: The total energy flux $dH$ into the system is composed of the 
575: \bfi{thermal energy flux} or \bfi{heat flux} $TdS$,
576: the \bfi{mechanical energy flux} $-PdV$, and the 
577: \bfi{chemical energy flux} $\mu \cdot dN$. 
578: 
579: The Gibbs-Duhem equation \gzit{e.GDp} describes the energy balance 
580: necessary to compensate the changes $d(TS)=TdS+SdT$ of
581: thermal energy, $d(PV)=Pd V + V dP$ of
582: mechanical energy, and $d(\mu \cdot N)=\mu \cdot dN+N\cdot d\mu$ 
583: of chemical energy in the energy contributions to the Euler equation 
584: to ensure that the Euler equation
585: remains valid during a reversible transformation. Indeed, both
586: equations together imply that $d(TdS-PdV+\mu\cdot N -H)$
587: vanishes, which expresses the preservation of the Euler equation.
588: 
589: Related to the various energy fluxes are the \bfi{thermal work}
590: \[
591: Q = \int T(\lambda)dS(\lambda),
592: \]
593: the \bfi{mechanical work}
594: \[
595: W_\fns{mech} = -\int P(\lambda)dV(\lambda),
596: \]
597: and the \bfi{chemical work}
598: \[
599: W_\fns{chem} = \int \mu(\lambda)\cdot dN(\lambda)
600: \]
601: performed in a reversible transformation. The various kinds of work 
602: generally depend on the path through the state space; however, the
603: mechanical work depends only on the end points if the associated
604: process is conservative. 
605: 
606: As is apparent from the formulas given, thermal work is done by
607: changing the entropy of the system, mechanical work by changing the 
608: volume, and chemical work by changing the mole numbers. 
609: In particular, in case of thermal, mechanical, or chemical
610: \bfi{isolation}, the corresponding fluxes vanish identically. 
611: Thus, constant $S$ characterizes thermally isolated, 
612: \bfi{adiabatic} systems, constant $V$ characterizes mechanically 
613: isolated, \bfi{closed}\footnote{
614: Note that the terms 'closed system' has also a 
615: much more general interpretation -- which we do {\em not} use in this 
616: chapter --, namely as a conservative dynamical system.
617: } % end footnote
618: systems, and constant $N$ 
619: characterizes chemically isolated, \bfi{impermeable} systems. 
620: Note that this constancy only holds when all assumptions for a 
621: standard system are valid: global equilibrium, a single phase, and 
622: the absence of chemical reactions.
623: Of course, these boundary conditions are somewhat idealized situations, 
624: which, however, can be approximately realized in practice and are of 
625: immense scientific and technological importance.
626: 
627: The first law shows that, in appropriate units, the temperature $T$ is 
628: the amount of energy needed to increase in a mechanically and 
629: chemically isolated system the entropy $S$ by one unit. 
630: The pressure $P$ is, in appropriate units, the amount of energy needed 
631: to decrease in a thermally and chemically isolated system the volume 
632: $V$ by one unit. In particular, increasing pressure decreases the 
633: volume; this explains the minus sign in the definition of $P$.
634: The chemical potential $\mu_j$ is, in appropriate units, the amount of 
635: energy needed to increase in a thermally and mechanically isolated 
636: system the mole number $N_j$ by one. With the traditional units, 
637: temperature, pressure, and chemical potentials are no longer energies.
638: 
639: We see that the entropy and the volume behave just like the mole
640: number. This analogy can be deepened by observing that mole numbers 
641: are the natural measure of the amounts of ''matter'' of each kind in a 
642: system, and chemical energy flux is accompanied by adding or removing 
643: matter.
644: Similarly, volume is the natural measure of the amount of ''space'' a
645: system occupies, and mechanical energy flux in a standard system is
646: accompanied by adding or removing space.
647: Thus we may regard entropy as the natural measure of the amount of 
648: ''heat'' contained in a system\footnote{
649: Thus, entropy is the modern replacement for the historical concepts of 
650: \bfi{phlogiston} and \bfi{caloric}, which failed to give a correct 
651: account of heat phenomena. 
652: Phlogiston turned out to be 'missing oxygen', an early 
653: analogue of the picture of positrons as holes, or 'missing electrons', 
654: in the Dirac sea. Caloric was a massless substance of heat which had 
655: almost the right properties, explained many effects correctly, and 
656: fell out of favor only after it became known that caloric could be
657: generated in arbitrarily large amounts from mechanical energy, thus
658: discrediting the idea of heat being a substance. (For the precise 
659: relation of entropy and caloric, see \sca{Kuhn} \cite{Kuh1,Kuh2},
660: \sca{Walter} \cite{Walt}, and the references quoted there.) 
661: In the modern picture,
662: the extensivity of entropy models the substance-like properties of
663: heat. But as there are no particles 
664: of space whose mole number is proportional to the volume, so there are 
665: no particles of heat whose mole number is proportional to the entropy.
666: Nevertheless, the introduction of heat particles on a formal level 
667: has some uses; see, e.g., \sca{Streater} \cite{Str}.
668: },  % end footnote
669: since thermal energy flux is accompanied by adding or removing heat.
670: Looking at other extensive quantities, we also recognize energy as the 
671: natural measure of the amount of ''power'', 
672: momentum as the natural measure of the amount of ''force'', and 
673: mass as the natural measure of the amount of ''inertia'' of a system. 
674: In each case, the notions in quotation marks are the colloquial terms 
675: which are associated in ordinary life with the more precise, formally 
676: defined physical quantities. For historical reasons, the words heat, 
677: power, and force are used in physics with a meaning different from the 
678: colloquial terms ''heat'', ''power'', and ''force''.
679: 
680: 
681: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
682: \section{Consequences of the second law}\label{s.c2}
683: 
684: The second law is centered around the impossibility of perpetual 
685: motion machines due to the inevitable loss of energy by  
686: dissipation such as friction (see, e.g., \sca{Bowden \& Leben} 
687: \cite{BowL}), uncontrolled radiation, etc..
688: This means that -- unless continually provided from the outside --
689: energy is lost with time until a metastable state is attained,
690: which usually is an equilibrium state. Therefore, the energy at
691: equilibrium is minimal under the circumstances dictated by the 
692: boundary conditions. In a purely kinematic setting as in our treatment, 
693: the approach to equilibrium cannot be studied, and only the  
694: minimal energy principles -- one for each set of boundary conditions -- 
695: remain. 
696: 
697: Traditionally, the second law is often expressed in the form of
698: an extremal principle for some thermodynamic potential.
699: We derive here the extremal principles for the Hamilton energy,
700: the Helmholtz energy, and the Gibbs energy\footnote{
701: The different potentials are related by so-called Legendre transforms;
702: cf. \sca{Rockafellar} \cite{Roc} for the mathematical properties of 
703: Legendre transforms, \sca{Arnol'd} \cite{Arn} for their application 
704: in mechanics, and \sca{Alberty} \cite{Alb} for their application in 
705: chemistry.
706: }, % end footnote
707: which give rise to the \bfi{Hamilton potential}
708: \[
709: U(S,V,N) :=\max_{T,P,\mu}\,
710: \{TS-PV+\mu\cdot N\mid \Delta(T,P,\mu)=0;T>0;P>0\},
711: \]
712: the \bfi{Helmholtz potential}
713: \[
714: A(T,V,N):=\max_{P,\mu}\,
715: \{-PV+\mu\cdot N\mid \Delta(T,P,\mu)=0;T>0;P>0\},
716: \]
717: and the \bfi{Gibbs potential}
718: \[ 
719: G(T,P,N):=\max_\mu\,
720: \{\mu\cdot N\mid \Delta(T,P,\mu)=0;T>0;P>0\}.
721: \]
722: The Gibbs potential is of particular importance for everyday 
723: processes since the latter frequently happen at approximately constant 
724: temperature, pressure, and mole number. 
725: (For other thermodynamic potentials used in practice, see 
726: \sca{Alberty} \cite{Alb}; for the maximum entropy 
727: principle, see Section \ref{s.maxent}.)
728:  
729: 
730: \begin{thm}\label{t.extstd}
731: \bfi{(Extremal principles)}\\
732: (i) In an arbitrary state, 
733: \lbeq{e.2ndU}
734: H \ge U(S,V,N),
735: \eeq
736: with equality iff the state is an equilibrium state. 
737: The remaining thermodynamic variables are then given by
738: \[
739: T = \frac{\partial}{\partial S}U(S,V,N),~~~
740: P = -\frac{\partial}{\partial V}U(S,V,N),~~~
741: \mu = \frac{\partial}{\partial N}U(S,V,N),~~~
742: H = U(S,V,N).
743: \]
744: In particular, an equilibrium state is uniquely determined by 
745: the values of $S$, $V$, and $N$.
746: 
747: (ii) In an arbitrary state, 
748: \lbeq{e.2ndA}
749: H-TS \ge A(T,V,N),
750: \eeq
751: with equality iff the state is an equilibrium state.
752: The remaining thermodynamic variables are then given by
753: \[
754: S=-\frac{\partial A}{\partial T}(T,V,N),~~~
755: P=-\frac{\partial A}{\partial V}(T,V,N),~~~
756: \mu=\frac{\partial A}{\partial N}(T,V,N),
757: \]
758: \[
759: H=A(T,V,N)+TS.
760: \]
761: In particular, an equilibrium state is uniquely determined by 
762: the values of $T$, $V$, and $N$.
763: 
764: (iii) In an arbitrary state,
765: \lbeq{e.2ndG}
766: H-TS+PV \ge G(T,P,N),
767: \eeq
768: with equality iff the state is an equilibrium state.
769: The remaining thermodynamic variables are then given by
770: \[
771: S=-\frac{\partial G}{\partial T}(T,P,N),~~~
772: V=\frac{\partial G}{\partial P}(T,P,N), ~~~
773: \mu=\frac{\partial G}{\partial N}(T,P,N),
774: \]
775: \[
776: H=G(T,P,N)+TS-PV.
777: \]
778: In particular, an equilibrium state is uniquely determined by 
779: the values of $T$, $P$, and $N$.
780: \end{thm}
781: 
782: \bepf
783: We prove (ii); the other two cases are entirely similar.
784: \gzit{e.2ndA} and the statement about equality is a direct consequnce 
785: of Axiom \ref{d.phen}(ii). Thus, the difference $H-TS-A(T,V,N)$ 
786: takes its minimum value zero at the equilibrium value of $T$. 
787: Therefore, the derivative with respect to $T$ vanishes, which gives the
788: formula for $S$. To get the formulas for $P$ and $\mu$, we note that 
789: for constant $T$, the first law \gzit{3.1stp} implies
790: \[
791: dA=d(H-TS)=dH-TdS=-PdV+\mu\cdot dN.
792: \]
793: For the reversible transformation which only changes $P$ or $\mu_j$,
794: we conclude that $dA=-PdV$ and $dA=\mu\cdot dN$, respectively.
795: Solving for $P$ and $\mu_j$, respectively, implies the formulas for
796: $P$ and $\mu_j$.
797: \epf
798: 
799: The above results imply that one can regard each thermodynamic 
800: potential as a complete alternative way to describe the manifold of 
801: thermal states and hence all equilibrium properties.
802: This is very important in practice, where one usually describes
803: thermodynamic material properties in terms of the Helmholtz or Gibbs 
804: potential, using models like NRTL (\sca{Renon \& Prausnitz} \cite{RenP},
805: \sca{Prausnitz} et al. \cite{PraLA})
806: or SAFT (\sca{Chapman} et al. \cite{ChaGJR,ChaGJR2}).
807: 
808: The additivity of extensive quantities is reflected in corresponding 
809: properties of the thermodynamic potentials:
810: 
811: \begin{thm}\label{t.ext}
812: The potentials $U(S,V,N)$, $A(T,V,N)$, and $G(T,P,N)$
813: satisfy, for real $\lambda,\lambda^1,\lambda^2\ge 0$,
814: \lbeq{e.homUx}
815: U(\lambda S,\lambda V,\lambda N)=\lambda U(S,V,N),
816: \eeq
817: \lbeq{e.homAx}
818: A(T,\lambda V,\lambda N)=\lambda A(T,V,N),
819: \eeq
820: \lbeq{e.homGx}
821: G(T,P,\lambda N)=\lambda G(T,P,N),
822: \eeq
823: \lbeq{e.convUx}
824: U(\lambda^1 S^1+\lambda^2S^2,\lambda^1 V^1+\lambda^2V^2,
825: \lambda^1 N^1+\lambda^2N^2)
826: \le \lambda^1 U(S^1,V^1,N^1)+\lambda^2 U(S^2,V^2,N^2),
827: \eeq
828: \lbeq{e.convAx}
829: A(T,\lambda^1 V^1+\lambda^2V^2,\lambda^1 N^1+\lambda^2N^2)
830: \le \lambda^1 A(T,V^1,N^1)+\lambda^2 A(T,V^2,N^2),
831: \eeq
832: \lbeq{e.convGx}
833: G(T,P,\lambda^1 N^1+\lambda^2N^2)
834: \le \lambda^1 G(T,P,N^1)+\lambda^2 G(T,P,N^2).
835: \eeq
836: In particular, these potentials are convex in $S$, $V$, and $N$.
837: \end{thm}
838: 
839: \bepf
840: The first three equations express homogeneity and are a direct 
841: consequence of the definitions. Inequality \gzit{e.convAx} holds since, 
842: for suitable $P$ and $\mu$,
843: \[
844: \bary{lll}
845: A(T,\lambda^1 V^1+\lambda^2V^2,\lambda^1 N^1+\lambda^2N^2)
846: &=&-P(\lambda^1 V^1+\lambda^2V^2)+\mu\cdot(\lambda^1 N^1+\lambda^2N^2)\\
847: &=&\lambda^1(-PV^1+\mu\cdot N^1)+\lambda^2(-PV^2+\mu\cdot N^2)\\
848: &\le& \lambda^1 A(T,V^1,N^1)+\lambda^2 A(T,V^2,N^2);
849: \eary
850: \]
851: and the others follow in the same way.
852: Specialized to $\lambda^1+\lambda^2=1$, the inequalities express the
853: claimed convexity.
854: \epf
855: 
856: For a system at constant temperature $T$, pressure $P$, and mole
857: number $N$, consisting of
858: a number of parts labeled by a superscript $k$ which are separately
859: in equilibrium, the Gibbs energy is extensive, since
860: \[
861: \bary{lll}
862: G&=&H-TS+PV= \D\sum H^k-T\sum S^k+P\sum V^k \\
863: &=& \D\sum (H^k-TS^k+PV^k)=\sum G^k.
864: \eary
865: \]
866: Equilibrium requires that $\sum G^k$ is minimal among all choices 
867: with $\sum N^k=N$, and by introducing a Lagrange multiplier vector 
868: $\mu^*$ for the constraints, we see that in equilibrium, the derivative 
869: of $\sum (G(T,P,N^k)-\mu^*\cdot N^k)$ with respect to each $N^k$ 
870: must vanish. This implies that 
871: \[
872: \mu^k= \frac{\partial G}{\partial N^k}(T,P,N^k)=\mu^*.
873: \]
874: Thus, in equilibrium, all $\mu^k$ must be the same.  
875: At constant $T$, $V$, and $N$, one can apply the same argument to the 
876: Helmholtz potential, and at constant $S$, $V$, and $N$ to the 
877: Hamilton potential. In each case, the equilibrium is characterized
878: by the constancy of the intensive parameters.
879: 
880: The degree to which 
881: macroscopic space and time correlations are absent characterizes
882: the amount of \bfi{macroscopic disorder} of a system.
883: Global equilibrium states are therefore macroscopically highly uniform; 
884: they are the most ordered macroscopic states in the universe rather 
885: than the most disordered ones. 
886: A system not in global equilibrium is characterized by macroscopic 
887: local inhomogeneities, indicating that the space-independent global 
888: equilibrium variables alone are not sufficient to describe the system.
889: Its intrinsic complexity is apparent only in a microscopic treatment;
890: cf. Section \ref{s.complexity} below.
891: The only macroscopic shadow of this complexity is the critical
892: opalescence of fluids near a critical point (\sca{Andrews} \cite{And}, 
893: \sca{Forster} \cite{For}). The contents of the second law of
894: thermodynamics for global equilibrium states may therefore be phrased 
895: informally as follows:
896: {\em In global equilibrium, macroscopic order (homogeneity) is perfect 
897: and microscopic complexity is maximal}.
898: In particular, the traditional interpretation of entropy
899: as a measure of disorder is often misleading.
900: Much more carefully argued support for this statement, with numerous
901: examples from teaching practice, is in \sca{Lambert} \cite{Lam}.
902: 
903: \begin{thm} \label{4.1.} (\bfi{Entropy form of the second law})\\
904: In an arbitrary state of a standard thermodynamic system 
905: \[
906: S \le S(H,V,N)
907: :=\min\,\{T^{-1}(H+PV-\mu\cdot N)\mid \Delta(T,P,\mu)=0\},
908: \]
909: with equality iff the state is an equilibrium state.
910: The remaining thermal variables are then given by
911: \lbeq{e.ent1x}
912: T^{-1}=\frac{\partial S}{\partial H}(H,V,N),~~~
913: T^{-1}P=\frac{\partial S}{\partial V}(H,V,N),~~~
914: T^{-1}\mu=-\frac{\partial S}{\partial N}(H,V,N),
915: \eeq
916: \lbeq{e.ent2x}
917: U=H=TS(T,V,N)-PV+\mu\cdot N.
918: \eeq
919: \end{thm}
920: 
921: \bepf
922: This is proved in the same way as Theorem \ref{t.extstd}.
923: \epf
924: 
925: This result implies that when a system in which $H$, $V$ and $N$ are 
926: kept constant reaches equilibrium, the entropy must have increaed.
927: Unfortunately, the assumption of constant $H$, $V$ and $N$
928: is unrealistic; such constraints are not easily realized in nature.
929: Under different constraints\footnote{
930: For example, if one pours milk into a cup of coffee, stirring mixes
931: coffee and milk, thus increasing complexity. Macroscopic order is
932: restored after some time when this increased complexity has become
933: macroscopically inaccessible. Since $T,P$ and $N$ are constant, 
934: the cup of coffee ends up in a 
935: state of minimal Gibbs energy, and not in a state of maximal entropy!
936: More formally, the first law shows that, for standard systems at fixed 
937: value of the mole number, the value of the entropy decreases 
938: when $H$ or $V$ 
939: (or both) decrease reversibly; this shows that the value of the entropy 
940: may well decrease if accompanied by a corresponding decrease of 
941: $H$ or $V$. The same holds out of equilibrium (though our 
942: equilibrium argument no longer applies); for example, the reaction 
943: 2 H${}_2$ $+$ O${}_2$ $\to$ 2 H${}_2$O (if catalyzed) 
944: may happen spontaneously at constant $T=25\,^\circ$C and $P=1$~atm, 
945: though it decreases the entropy. 
946: }, % end footnote
947: the entropy is no longer maximal. 
948: 
949: In systems with several phases, a naive interpretation of the second 
950: law as moving systems towards increasing disorder is even more 
951: inappropriate: 
952: A mixture of water and oil spontaneously separates, thus ''ordering'' 
953: the water molecules and the oil molecules into separate phases! 
954: 
955: Thus, while the second law in the 
956: form of a maximum principle for the entropy has some theoretical and 
957: historical relevance, it is not the extremal principle ruling nature.
958: The irreversible nature of physical processes is instead manifest as 
959: \bfi{energy dissipation} which, in a microscopic interpretation, 
960: indicates 
961: the loss of energy to the unmodelled microscopic degrees of freedom.
962: Macroscopically, the global equilibrium states are therefore states 
963: of least free energy, the correct choice of which depends on the 
964: boundary condition, with the least possible freedom for change. 
965: This macroscopic immutability is another intuitive explanation for the 
966: maximal macroscopic order in global equilibrium states.
967: 
968: 
969: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
970: \section{The approach to equilibrium}\label{s.appeq}
971: 
972: Using only the present axioms, one can say a little bit about the
973: behavior of a system close to equilibrium in the following,
974: idealized situation.
975: Suppose that a system at constant $S$, $V$, and $N$ which is close to 
976: equilibrium at some time $t$ reaches equilibrium at some later time 
977: $t^*$. Then the second law implies 
978: \[
979: 0\le H(t)-H(t^*) \approx (t-t^*)\frac{dH}{dt},
980: \]
981: so that $dH/dt\le 0$. We assume that the system is composed of two 
982: parts, which are both in equilibrium at times $t$ and $t^*$. Then
983: the time shift induces on both parts a reversible transformation, 
984: and the first law can be applied to them. Thus
985: \[
986: dH=\sum_{k=1,2} dH^k =\sum_{k=1,2} (T^kdS^k-P^kdV^k+\mu^k\cdot dN^k).
987: \]
988: Since $S$, $V$, and $N$ remain constant, we have $dS^1+dS^2=0$,
989: $dV^1+dV^2=0$, $dN^1+dN^2=0$, and since for the time shift $dH\le 0$,
990: we find the inequality
991: \[
992: 0\ge (T^1-T^2)dS^1 - (P^1-P^2)dV^1 +(\mu^1-\mu^2)\cdot dN^1.
993: \]
994: This inequality gives infromation about the direction of the flow 
995: in case that all but one of the extensive variables are known to be 
996: fixed.
997: 
998: In particular, at constant $V^1$ and $N^1$, we have $dS^1\le 0$ if 
999: $T^1>T^2$; i.e., ''heat'' (entropy) flows from the hotter part towards 
1000: the colder part. At constant $S^1$ and $N^1$, we have $dV^1\le 0$ if 
1001: $P^1<P^2$; i.e., ''space'' (volume) flows from lower pressure to 
1002: higher pressure: the volume of the lower pressure part decreases and 
1003: is compensated by a corresponding increase of the volume in the higher 
1004: pressure part. And for a pure substance at constant $S^1$ and 
1005: $V^1$, we have $dN^1\le 0$ if $\mu^1>\mu^2$; i.e., ''matter'' (mole 
1006: number) flows from higher chemical potential towards lower chemical 
1007: potential. These qualitative results give temperature, pressure,
1008: and chemical potential the familiar intuitive interpretation. 
1009: 
1010: This glimpse on nonequilibrium properties is a shadow of the far 
1011: reaching fact that, in nonequilibrium 
1012: thermodynamics, the intensive variables behave like potentials whose
1013: gradients induce forces that tend to diminish these gradients,
1014: thus enforcing (after the time needed to reach equilibrium) agreement 
1015: of the intensive variables of different parts of a system. 
1016: In particular, temperature acts as a thermal potential, whose 
1017: differences create thermal forces which induce thermal currents, 
1018: a flow of ''heat'' (entropy), in a similar way as differences in 
1019: electrical potentials create electrical currents, a flow of 
1020: ''electricity'' (electrons)\footnote{
1021: See Table \ref{3.t.} for more parallels in other thermodynamic systems,
1022: and \sca{Fuchs} \cite{Fuc} for a thermodynamics course (and for a 
1023: German course \sca{Job} \cite{Job}) 
1024: thoroughly exploiting these parallels. 
1025: }. % end footnote
1026: While these dynamical issues are outside the scope 
1027: of the present work, they motivate the fact that one can control some
1028: intensive parameters of the system by controlling the corresponding 
1029: intensive parameters of the environment and making the walls permeable 
1030: to the corresponding extensive quantities. This
1031: corresponds to standard procedures familiar to everyone from ordinary 
1032: life, such as: heating to change the temperature; applying pressure 
1033: to change the volume; immersion into a substance to change the chemical 
1034: composition; or, in the more general thermal models discussed in 
1035: Section \ref{s.detail}, applying forces to displace an object.
1036: 
1037: The stronger nonequilibrium version of the second law says that 
1038: (for suitable boundary conditions) equilibrium is actually attained 
1039: after some time (stictly speaking, only in the limit of infinite time).
1040: This implies that the energy difference 
1041: \[
1042: \delta E:=H-U(S,V,N)=H-TS-A(S,V,N)=H-TS+PV=G(S,V,N)
1043: \]
1044: is the amount of energy that is dissipated in order to reach 
1045: equilibrium. In an equilibrium setting, we 
1046: can only compare what happens to a system prepared in a nonequilibrium 
1047: state assuming that, subsequently, the full energy difference 
1048: $\delta E$ is dissipated so that the system ends up in an equilibrium 
1049: state. Since few variables describe everything of interest, this 
1050: constitutes the power of equilibrium thermodynamics. But this power is 
1051: limited, since equilibrium thermodynamics is silent about when -- or 
1052: whether at all -- equilibrium is reached. Indeed, in many cases, only 
1053: metastable states are reached, which change too slowly to ever reach 
1054: equilibrium on a human time scale.
1055: 
1056: 
1057: \bigskip
1058: \bfi{Description levels.}
1059: As we have seen, extensive and intensive variables play completely 
1060: different roles in equilibrium thermodynamics. Extensive variables 
1061: such as mass, charge, or volume depend additively on the size of the 
1062: system. The conjugate intensive variables act as parameters defining 
1063: the state. 
1064: 
1065: A system composed of many small subsystems, each in equilibrium,
1066: needs for its complete characterization the values of the extensive and
1067: intensive variables in each subsystem. Such a system
1068: is in global equilibrium only if its intensive variables are independent
1069: of the subsystem. On the other hand, the values of the extensive
1070: variables may jump at phase space beoundaries, if (as is the case for 
1071: multi-phase systems) the equations of state allow multiple values for 
1072: the extensive variables to correspond to the same values of the 
1073: intensive variables.
1074: If the intensive variables are not independent of the subsystem then,
1075: by the second law, the differences in the intensive variables of 
1076: adjacent subsystems give rise to thermodynamic forces trying to move 
1077: the system towards equilibrium. 
1078: 
1079: A real nonequilibrium system does not actually consist of subsystems in
1080: equilibrium; however, typically, smaller and smaller pieces behave 
1081: more and more like equilibrium systems. Thus we may view a real system
1082: as the continuum limit of a larger and larger number of smaller and 
1083: smaller subsystems, each in approximate equilibrium. As a result, the 
1084: extensive and intensive variables become fields depending on the 
1085: continuum variables used to label the subsystems. 
1086: For extensive 
1087: variables, the integral of their fields over the label space gives 
1088: the bulk value of the extensive quantity; thus the fields themselves 
1089: have a natural interpretation as a density. For intensive variables,
1090: an interpretation as a density is physically meaningless; instead,
1091: they have a natural interpretation as field strengths. 
1092: The gradients of their fields have physical significance as the 
1093: sources for thermodynamic forces. 
1094: 
1095: From this fied theory perspective, the extensive variables in the 
1096: single-phase global equilibrium case have constant densities, and 
1097: their bulk values are the densities multiplied by the 
1098: system size (which might be mass, or volume, or another additive 
1099: parameter), hence scale linearly with the size of the system, while 
1100: intensive variables are invariant under a change of system size. 
1101: We do {\em not} use the alternative convention to call extensive any
1102: variable that scales linearly with the system size, and intensive
1103: any variable that is invariant under a change of system size.
1104: 
1105: We distinguish four nested levels of 
1106: thermal descriptions, depending on whether the system is considered 
1107: to be in global, local, microlocal, or quantum 
1108: equilibrium. The highest and computationally simplest level, 
1109: \bfi{global equilibrium}, is concerned with macroscopic situations 
1110: characterized by finitely many space- and time-independent variables. 
1111: The next level, \bfi{local equilibrium}, treats macroscopic situations 
1112: in a continuum mechanical description, where the equilibrium 
1113: subsystems are labeled by the space coordinates. Therefore the relevant 
1114: variables are finitely many space- and time-dependent fields.
1115: The next deeper level, \bfi{microlocal}\footnote{
1116: The term microlocal for a phase space dependent analysis is taken
1117: from the literature on partial differential equations; see, e.g., 
1118: \sca{Martinez} \cite{Mar}.
1119: }~ % end footnote
1120: \bfi{equilibrium}, treats mesoscopic situations 
1121: in a kinetic description, where the equilibrium subsystems are labeled 
1122: by phase space coordinates. The relevant variables are now finitely 
1123: many fields depending on time, position, and momentum; 
1124: cf. \sca{Balian} \cite{Bal2}. The bottom level is the 
1125: microscopic regime, where we must consider \bfi{quantum equilibrium}.
1126: This no longer fits a thermodynamic framework but must be described 
1127: in terms of quantum dynamical semigroups; see Section \ref{s.model}.
1128: 
1129: The relations between the different description levels are
1130: discussed in Section \ref{s.model}. Apart from descriptions on these 
1131: clear-cut levels, there are also various hybrid descriptions, where 
1132: some part of a system is described on a more detailed 
1133: level than the remaining parts, or where, as for stirred chemical 
1134: reactions, the fields are considered to be spatially homogeneous and 
1135: only the time-dependence matters. 
1136: 
1137: In global equilibrium, all thermal variables are constant 
1138: throughout the system, except at phase boundaries, where the extensive 
1139: variables may exhibit jumps and only the intensive variables remain 
1140: constant. This is sometimes referred to as the zeroth law of 
1141: thermodynamics and characterizes global equilibrium; it allows one to
1142: measure intensive variables (like temperature) by bringing a
1143: calibrated instrument that is sensitive to this variable
1144: (for temperature a thermometer) into equilibrium with the system to be 
1145: measured. For local or microlocal equilibrium, the same intuition 
1146: applies, but with fields in place of variables. Then extensive 
1147: variables are densities represented by distributions that can be 
1148: meaningfully integrated over bounded regions, whereas intensive 
1149: variables are nonsingular fields (e.g., pressure) whose integrals 
1150: are physically irrelevant.
1151: 
1152: 
1153: 
1154: 
1155: