0811.0222/ig.tex
1: \documentclass[preprint,showpacs,aps]{revtex4}
2: \usepackage{graphicx}% Include figure files
3: %\usepackage{dcolumn}% Align table columns on decimal point
4: %\usepackage{bm}% bold math
5: \begin{document}
6: \def\be{\begin{equation}}
7: \def\ee{\end{equation}}
8: \def\bea{\begin{eqnarray}}
9: \def\eea{\end{eqnarray}}
10: \title{Invariant geometry of the ideal gas}
11: % use optional labels to link authors explicitly to addresses:
12: % \author[label1,label2]{}
13: % \address[label1]{}
14: % \address[label2]{}
15: \author{Hernando Quevedo}
16: \email{quevedo@nucleares.unam.mx}
17: \affiliation{Instituto de Ciencias Nucleares, 
18: Universidad Nacional Aut\'onoma de M\'exico  \\
19:  AP 70543, M\'exico, DF 04510, MEXICO}
20:  \author{Alberto S\'anchez}
21: \email{asanchez@nucleares.unam.mx}
22: \affiliation{Instituto de Ciencias Nucleares, 
23: Universidad Nacional Aut\'onoma de M\'exico  \\
24:  AP 70543, M\'exico, DF 04510, MEXICO}
25:  \author{Alejandro V\'azquez}
26: \email{alec_vf@nucleares.unam.mx}
27: \affiliation{Facultad de Ciencias, 
28: Universidad Aut\'onoma del Estado de Morelos \\
29: Av. Universidad 1001,   
30:  Cuernavaca, MO 62210 MEXICO}
31: \begin{abstract}
32: % Text of abstract
33: 
34: We analyze a Legendre invariant 
35: metric structure in the space of thermodynamic
36: equilibrium states of an ideal gas. Due to the lack of 
37: thermodynamic interaction, the geometry 
38: turns out to be flat.  
39: We introduce the concept of thermodynamic geodesics,
40: which correspond to quasi-static processes, 
41: analyze their properties,  and show
42: that they can be used to determine the ``arrow of time" 
43: and to split the equilibrium space of the ideal gas into two completely 
44: different regions, separated by adiabatic geodesics which correspond to 
45: reversible thermodynamic processes.
46: 
47: 
48: 
49: 
50: \end{abstract}
51: \pacs{05.70.-a, 02.40.-k}
52: 
53: \maketitle
54: %\begin{keyword}
55: % keywords here, in the form: keyword \sep keyword
56: % Equilibrium thermodynamics \sep contact structure \sep Riemannian structure \sep 
57: % Weinhold's metric
58: % PACS codes here, in the form: \PACS code \sep code
59: %\end{keyword}
60: 
61: % main text
62: \section{Introduction}
63: \label{sec:int}
64: % Differential geometry is a very important tool of modern science, 
65: % specially of mathematical physics and its applications in physics,
66: % chemistry and engineering. 
67: The idea of using differential geometry in thermodynamics is due to Gibbs \cite{gibbs}
68: who realized that the first law of thermodynamics can be represented in terms of
69: what today is known as differential forms. Charatheodory \cite{car} interpreted 
70: the laws of thermodynamics in an axiomatic way and in terms of Pfaffian forms. 
71: Later on, Hermann \cite{her} introduced the concept of thermodynamic phase space where 
72: the thermodynamic variables play the role of coordinates. Additionally, Weinhold 
73: \cite{wei75} and Ruppeiner \cite{rup79,rup95} introduced  metric structures in the
74: space of equilibrium states. The connection between the structure of the phase space
75: and the equilibrium space was recently incorporated \cite{quev07} in the formalism 
76: of {\it geometrothermodynamics} (GTD) in an invariant manner. In fact, a problem of 
77: using Weinhold's or Ruppeiner's metrics in equilibrium space is that the results
78: can depend on the choice of thermodynamic potential, i. e., the results are not invariant
79: with respect to Legendre transformations. GTD incorporates Legendre invariance into the
80: geometric structures of the phase space and equilibrium space so that the results do not
81: depend on the choice of thermodynamic potential or representation, a property which 
82: characterizes ordinary thermodynamics. 
83: 
84: One of the main results of GTD is that for any thermodynamic system it delivers a Legendre 
85: invariant metric $g=g_{ab}dx^a dx^b$ for the space of equilibrium states ${\cal E}$. 
86: Here $x^a$ are the coordinates in ${\cal E}$ and the components of the metric are in general 
87: functions of the coordinates, i. e. $g_{ab}=g_{ab} (x^c)$. In this manner, the equilibrium 
88: space becomes a Riemannian manifold whose geometric properties should be related to the properties
89: of the corresponding thermodynamic system. The most important geometric objects of a Riemannian 
90: manifold are the Christoffel symbols (or connection components)
91: \be
92: \label{symbols}
93: \Gamma^a_{\ bc} = \frac{1}{2} g^{ad}\left( \frac{\partial g_{db}}{\partial x^c}
94: +\frac{\partial g_{cd}}{\partial x^b}
95: -\frac{\partial g_{bc}}{\partial x^d}\right)\ ,
96: \ee
97: and the curvature tensor 
98: \be 
99: \label{curvature}
100: R^a_{\ bcd} = \frac{\partial \Gamma^a_{\ bd}}{\partial x^c} - \frac{\partial \Gamma^a_{\ bc}}{\partial x^d} +
101: \Gamma^a_{\ e c}\Gamma^e_{\ bd} - \Gamma^a_{\ e d}\Gamma^e_{\ bc} \ ,
102: \ee
103: which will be used in this work to investigate the properties of the equilibrium space.
104: We assume the convention of summation over repeated indices. Furthermore, from 
105: the curvature tensor one can define the Ricci tensor $R_{ab} = g^{cd} R_{acbd}$
106: and the curvature scalar $R = g^{ab}R_{ab}$. In GTD, the curvature tensor is expected to be
107: a measure of the thermodynamic interaction. In fact, we will present a metric whose curvature 
108: tensor vanishes in the case of an ideal gas, where thermodynamic interaction is absent. The curvature
109: tensor is not zero for systems with intrinsic thermodynamic interaction. Because of this property, 
110: the curvature of the space of equilibrium is called {\it thermodynamic curvature}.
111: 
112:  
113: The Christoffel symbols are important in connection with geodesics which are extremal curves 
114: $x^a(\tau)$ satisfying the differential equations
115: \be
116: \label{geodesics}
117: \frac{d^2x^a}{d\tau^2} + \Gamma^a_{ \ bc} \frac{dx^b}{d\tau} \frac{dx^c}{d\tau} = 0 \ ,
118: \ee
119: where $\tau$ is an affine parameter along the geodesics. We introduce the concept of {\it thermodynamic 
120: geodesic} as those solutions of the geodesic equations which fulfill the laws of thermodynamics. 
121: We show that thermodynamic geodesics describe quasi-static processes and that $\tau$ can be used as
122: a ``time" parameter. Moreover, thermodynamic geodesics have a definite direction that can be interpreted 
123: as the ``arrow of time". 
124: 
125: The main goal of this work is to present a Legendre invariant metric for the space of equilibrium 
126: states of an ideal gas and investigate in detail its geodesics. Related investigations were carried
127: out in \cite{nulsal85,john03,san05c} by using, however, metric structures which depend on the choice of 
128: thermodynamic potential. On the contrary, we show explicitly that our results are independent of the choice 
129: of thermodynamic  potential and, therefore, represent an invariant representation of the geometry of the ideal gas.
130: 
131: This paper is organized as follows. In Section \ref{sec:gtdbh} we present the fundamentals of GTD in the case 
132: of systems with two thermodynamic degrees of freedom. In particular, we present a Legendre invariant metric for
133: the phase space from which it is possible to derive an invariant geometry for the space of
134: equilibrium states. In Section \ref{sec:ig} we focus on the geometry of the ideal gas and show that the
135: corresponding thermodynamic curvature vanishes as a result of the lack of thermodynamic interaction.
136: Furthermore, in Sections \ref{sec:geoig} and \ref{sec:geotv} we perform a detailed analysis of the geodesic
137: equations. The equilibrium space presents a very rich and unexpected structure that resembles the 
138: causality structure of spacetime in relativistic physics.
139: In Appendix \ref{sec:app1} we present some details of Legendre transformations and different metrics
140: that arise in the entropy representation of the ideal gas. 
141:  Finally, Section \ref{sec:con} is devoted
142: to discussions of our results and suggestions for further research.
143: % Throughout this paper we use units in which $G=c=k_{_B}=\hbar =1$.
144: 
145: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
146: \section{Geometrothermodynamics of simple systems}
147: \label{sec:gtdbh}
148: 
149: 
150: The simplest thermodynamic systems are those with only two thermodynamic degrees 
151: of freedom. To describe this type of systems, it is necessary to specify two 
152: extensive variables, say entropy $S$ and volume $V$, two intensive variables, 
153: say temperature $T$ and pressure $P$, and a thermodynamic potential, say the
154: internal energy $U$. In GTD \cite{quezar03,quev07}, the thermodynamic variables are used as 
155: coordinates for the construction of the thermodynamic phase space ${\cal T}$,
156: which in this case is 5-dimensional. We assume that ${\cal T}$ is a 
157: differential manifold so that it allows the introduction of additional 
158: geometric structures. In particular, we introduce the so called fundamental 
159: Gibbs form
160: \be
161: \label{gibbs}
162: \Theta = dU - T d S + P d V \ ,
163: \ee
164: which, as we will see below, contains the information about the first law of 
165: thermodynamics. In addition, we introduce a metric $G$ that, in
166: general, can depend on all the coordinates of ${\cal T}$, i.e. $G=G(U,S,V,T,P)$.
167: 
168: The triplet $({\cal T},\Theta, G)$ is called a Riemannian contact manifold 
169: and represents an auxiliary structure which is necessary to implement 
170: in a consistent manner the properties of ordinary thermodynamics. In particular,
171: it is known that the thermodynamic properties of a system do not depend on 
172: the thermodynamic potential used for its description \cite{callen}. 
173: Since different thermodynamic
174: potentials are related by means of Legendre transformations, we demand that 
175: the geometric structure of the phase space ${\cal T}$ be Legendre invariant. 
176: In the case of a system with two degrees of freedom, 
177: a Legendre transformation is defined in ${\cal T}$ 
178: as the change of coordinates $(U,S,V,T,P)\longrightarrow 
179: (\tilde U,\tilde S,\tilde V,\tilde T,\tilde P)$ with the following 
180: possibilities \cite{arnold}:
181: \bea
182: \label{lt1}
183: \tilde U_1 = U - TS \ ,\quad S = -\tilde T\ ,\quad T =\tilde S \ ,\quad V = \tilde V
184: \ ,\quad P =\tilde P \ ,\\
185: \label{lt2}
186: \tilde U_2 = U + PV \ ,\quad S = \tilde S\ ,\quad T =\tilde T \ ,\quad V = \tilde P
187: \ ,\quad - P =\tilde V \ ,\\
188: \label{lt3}
189: \tilde U_3 = U - TS + PV \ ,\quad S = -\tilde T\ ,\quad T =\tilde S \ ,\quad V = \tilde P
190: \ ,\quad - P =\tilde V \ .
191: \eea
192: Usually, $\tilde U_1=F$ is called the Helmholtz free energy, $\tilde U_2=H$ is 
193: the enthalpy, and $\tilde U_3=G$ is the Gibbs potential. If we denote by $\tilde \Theta_i$, 
194: $i=1,2,3$, the result of applying any of the particular Legendre transformations 
195: (\ref{lt1})-(\ref{lt3}) to the fundamental Gibbs form, then it is easy to see that
196: $\tilde \Theta_i = d\tilde U_i - \tilde T d \tilde S +\tilde P d\tilde V$, showing that
197: in fact $\Theta$ is a Legendre invariant geometric object. 
198: 
199: In GTD we also demand that the metric $G$ be Legendre invariant. It is possible to write
200: down and solve the algebraic conditions that an arbitrary metric $G$ must satisfy 
201: in order to be Legendre invariant \cite{quev07}. A particular solution was found in 
202: \cite{vqs08} that can be written as
203: \be
204: \label{ginv3}
205: G=\left(dU - T d S + P d V\right)^2 + \Lambda \left[(ST)^{2k+1} dS dT + (VP)^{2k+1} dV dP\right] \ ,
206: \ee
207: where $k$ is an integer, positive or negative, and 
208: $\Lambda=\Lambda(U,S,V,T,P)$ must be chosen as 
209: a Legendre invariant function of all its arguments. As in the case of the
210: fundamental Gibbs form, if we denote by $\tilde G_i$
211: the result of applying any of the Legendre transformations (\ref{lt1})-(\ref{lt3})
212: to the metric (\ref{ginv3}), the Legendre invariance of $G$ becomes clear. 
213: In fact, the functional dependence of $\tilde G_i$  coincides with (\ref{ginv3}) with 
214: $U$ replaced by $\tilde U_i$, $S$ by $\tilde S$, and so on. In this manner, we see that 
215: the particular triplet $({\cal T}, \Theta, G)$, with $\Theta$ and $G$ given as in Eqs.(\ref{gibbs}) and
216: (\ref{ginv3}), respectively, is invariant with respect to all possible Legendre transformations
217: in the case of systems with two thermodynamic degrees of freedom. This is an important property which 
218: guarantees that our further results are independent of the choice of thermodynamic potential.
219: 
220: The next important element of GTD is the space of equilibrium states ${\cal E}$ which, in quite
221: general terms, is the space where systems in thermodynamic equilibrium can exist and 
222: their properties can be 
223: investigated. This means that ${\cal E}$ is a 2-dimensional subspace of ${\cal T}$ which we define in the following 
224: manner. Let us choose the set of extensive variables $(S,V)$ as the coordinates of ${\cal E}$. Then, when evaluated 
225: on ${\cal E}$, the remaining coordinates of ${\cal T}$ must be functions of $U$ and $V$ only, i. e.
226: \be
227: \label{coor1}
228: U=U(S,V)\ ,\quad T=T(S,V)\ ,\quad P=P(S,V)\ .
229: \ee
230: The first of these equations is known as the fundamental equation in the energy representation. In fact, once 
231: $U(S,V)$ is given explicitly, one can derive all the equations of state and thermodynamic properties of the
232: corresponding thermodynamic system.  To guarantee the existence of the second and third equations of (\ref{coor1}), we
233: demand that the projection of the fundamental Gibbs form on ${\cal E}$ vanishes, i. e.,
234: \be
235: \label{firstlaw}
236: \Theta |_{\cal E} = 0 \Longleftrightarrow dU = T d S - P d V \ ,
237: \ee
238: a relationship that is immediately recognized as the first law of thermodynamics. Furthermore, since $U=U(S,V)$, the first
239: law of thermodynamics implies that
240: \be
241: \label{eqcond}
242: \frac{\partial U}{\partial S} = T\ ,\quad \frac{\partial U}{\partial V} = - P\ ,
243: \ee
244: so that $T$ and $P$ become functions of $S$ and $V$, as stated in (\ref{coor1}). In ordinary thermodynamics, the relationships
245: (\ref{eqcond})
246: represent the conditions for thermodynamic equilibrium. As for the metric $G$ of ${\cal T}$, we demand that its projection
247: on ${\cal E}$, by using (\ref{firstlaw}) and (\ref{eqcond}),   induces a metric $g$ on ${\cal E}$, i. e., $G|_{\cal E} = g = g(S,V)$.
248: In the particular case of the Legendre invariant metric (\ref{ginv3}), a straightforward calculation leads to
249: \bea
250: \label{gdowninv1}
251: g  = && \Lambda \Bigg\{ 
252: \left(S\frac{\partial U}{\partial S}\right)^{2k+1}
253: \frac{\partial^2 U}{\partial S ^2} d S^2
254: + \left(V\frac{\partial U}{\partial V}\right)^{2k+1} 
255: \frac{\partial^2 U}{\partial V ^2} d V^2 \\
256: & & +   \left[ \left(S\frac{\partial U}{\partial S}\right)^{2k+1}
257: +\left(V\frac{\partial U}{\partial V}\right)^{2k+1}\right]
258: \frac{\partial^2 U}{\partial S \partial V} d S d V  \Bigg\} \ .
259: \eea
260: If we fix the function $\Lambda$ and the constant $k$, and specify a fundamental equation $U=U(S,V)$, the above metric is unique. This means
261: that for  a given thermodynamic system, the metric (\ref{gdowninv1}) determines a family of geometries that characterizes 
262: the corresponding space of equilibrium ${\cal E}$. In fact, we will see that the freedom contained in the choice of $\Lambda$ and $k$ 
263: can be used to impose further physical conditions on the geometry of ${\cal E}$.
264: 
265: The above description corresponds to the energy representation in ordinary thermodynamics. One of the advantages of GTD is that the formalism
266: allows us to handle different representations in an invariant way. For later purposes, we present here the entropy representation for a system
267: with two thermodynamic degrees of freedom. A simple rearrangement of Eq.(\ref{gibbs}) leads to the Gibbs form 
268: in the entropy representation
269: \be
270: \Theta_S = dS - \frac{1}{T}dU - \frac{P}{T} dV \ ,
271: \ee
272: so that the coordinates of the phase space ${\cal T}$ are $(S,U,V,1/T,P/T)$, 
273: and the metric $G$ corresponding
274: to (\ref{ginv3}) can be written as
275: \be
276: \label{gups}
277: G_S = \left(dS -\frac{1}{T} d U - \frac{P}{T} dV\right)^2
278: + \Lambda \left[\left(\frac{U}{T}\right)^{2k+1} dU d\left(\frac{1}{T}\right) 
279: +\left(\frac{VP}{T}\right)^{2k+1} dV d\left(\frac{P}{T}\right)\right] \ .
280: \ee
281: With the corresponding change of coordinates (see Appendix \ref{sec:app1}), 
282: it is easy to show that the above geometric objects are invariant
283: with respect to Legendre transformations. Furthermore, for the equilibrium subspace ${\cal E}$ we choose the
284: extensive variables $U$ and $V$ so that the remaining coordinates become functions of $U$ and $V$ when projected 
285: on ${\cal E}$. In particular, the fundamental equation must now be given as $S=S(U,V)$. As before, we demand that
286: the projected Gibbs form and metric satisfy
287: \be
288: \Theta_S|_{{\cal E}}=0 \ , \quad G_S|_{{\cal E}} = g_S=g_S(S,V) \ ,
289: \ee
290: so that from the first condition we obtain the first law of thermodynamics and the conditions for thermodynamic
291: equilibrium in the entropy representation:
292: \be
293: \label{eqconds}
294: d S = \frac{1}{T}d U + \frac{P}{T} dV \ ,\quad
295: \frac{\partial S}{\partial U} = \frac{1}{T}\ , \quad \frac{\partial S}{\partial V} = \frac{P}{T}\ .
296: \ee
297: Moreover, the metric $g_S$ of the equilibrium space can be calculated in a straightforward manner from the above equations and we obtain
298: \bea
299: \label{gdowns}
300: g_S= \Lambda &\Bigg\{&\left(U\frac{\partial S}{\partial U}\right)^{2k+1}\frac{\partial^2 S}
301: {\partial U^2} dU^2
302: +  \left(V\frac{\partial S}{\partial V}\right)^{2k+1}
303: \frac{\partial^2 S}{\partial V^2} dV^2 \nonumber\\
304: & +&\left[ \left(U\frac{\partial S}{\partial U}\right)^{2k+1}
305: + \left(V\frac{\partial S}{\partial V}\right)^{2k+1} \right] 
306: \frac{\partial^2 S}{\partial U \partial V} dU dV \ \Bigg\} \ .
307: \eea
308: Again we see that once the fundamental equation $S=S(U,V)$ is given, the metric of ${\cal E}$ is uniquely determined, up to a constant $k$ 
309: and the conformal function $\Lambda$. As in ordinary thermodynamics, the fundamental equation must satisfy the second law which in the case 
310: under consideration can be written as \cite{callen} 
311: \be
312: \label{second}
313: \frac{\partial^2 U}{\partial E^a \partial E^b} \geq 0\ , \quad 
314: \frac{\partial^2 S}{\partial F^a \partial F^b} \leq 0 \ ,
315: \ee
316: for the energy and entropy representation, respectively. Here $a, b = 1,2\ ,  E^a = (S,V)$ and $F^a = (U,V)$. 
317: 
318: It is worth mentioning that in the above construction we chose  the extensive variables as coordinates for the equilibrium space 
319: ${\cal E}$ in order to obtain the energy and entropy representation which are the most common approaches used in ordinary thermodynamics. However, in general it is possible to choose any 2-dimensional subspace of ${\cal T}$ to define the
320:  equilibrium space ${\cal E}$ which would correspond to a different representation in ordinary thermodynamics. 
321:  The Legendre invariance of the phase space ${\cal T}$ has as a
322: consequence that all possible representations are equivalent and the properties of thermodynamic systems do not depend on the representation.
323: 
324: The above geometric construction can be applied to any thermodynamic system with two degrees of freedom. One only needs to specify the fundamental
325: equation of the thermodynamic system in order to investigate its geometric 
326:  properties. However, GTD allows a generalization to include any system with
327: a finite number of degrees of freedom, say $n$. In this case, the phase space has the dimension $2n+1$ and the subspace of equilibrium states is 
328: $n-$dimensional. The fundamental Gibbs form and the metrics $G$ and $g$ can be generalized to the $(2n+1)-$dimensional case  in a straightforward
329: manner \cite{vqs08}. 
330: 
331: 
332: 
333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335: \section{The equilibrium space of the ideal gas}
336: \label{sec:ig}
337: 
338: In the specific case of an ideal gas, the fundamental equation in the entropy representation 
339: can be expressed as \cite{callen}
340: \be
341: \label{feig}
342: S(U,V) = S_0 + N k_{_B} c_{_V}\ln\left(\frac{U}{ U_0}\right)  + N k_{_B} 
343: \ln\left(\frac{V}{ V_0}\right) \ ,
344: \ee
345: where $c_{_V}$ is the dimensionless 
346: heat capacity at constant volume, $k_{_B}$ is Boltzmann's constant, 
347: $N$ is the number of particles and $S_0$, $U_0$ 
348: and $V_0$ are constants. The intensive thermodynamic variables can be calculated by using the conditions
349: of thermodynamic equilibrium (\ref{eqconds}). We obtain
350: \be
351: \label{steq}
352: \frac{1}{T} = \frac{N k_{_B} c_{_V}}{U}\ ,\quad
353: \frac{P}{T} = \frac{N k_{_B}}{V}\ .
354: \ee
355: 
356: Furthermore, from Eq.(\ref{gdowns}) we obtain the simple metric 
357: \be
358: \label{gdownig}
359: g_S = -(N k_{_B})^{2k+2} \Lambda(U,V)\left[ c_{_V}^{2k+2}\frac{dU^2}{U^2} +  \frac{dV^2}{V^2}\right]\ .
360: \ee
361: The arbitrariness contained in the function $\Lambda$ and the constant $k$ can be fixed by demanding that 
362: $g_S$ determines an extremal surface \cite{burke}. 
363: This condition is very common in physics and is usually interpreted as
364: a consequence of the principle of minimum action. In fact, the metric (\ref{gdownig}) determines the infinitesimal area 
365: element $dA= \sqrt{|\det (g_S)|}\, d U d V$ so that the calculation of an arbitrary area in the space determined by $g_S$
366: implies the calculation of the integral $A = \int \sqrt{|\det (g_S)|}\, d U d V$. If we demand that the area be an extremal, 
367: i. e. $\delta A = 0$, where $\delta$ represents the variation, we obtain a set of differential equations \cite{vqs08}, implying
368: conditions on the form of the metric $g_S$. 
369: In the case of the metric (\ref{gdownig}) it can be shown that those differential equations are identically satisfied 
370: if we choose the particular solution $k=-1$ and $\Lambda = const$. For the sake of simplicity, let us choose $\Lambda=-1$
371: so that the metric (\ref{gdownig}) becomes
372: \be
373: \label{gdownig1}
374: g_S= \frac{dU^2}{U^2} +  \frac{dV^2}{V^2}
375: \ee
376: From this metric one can immediately calculate the corresponding connection and curvature, according to the formulas given in Section 
377: \ref{sec:int}, and show that the connection has non-vanishing components, but all the components of the thermodynamic 
378: curvature are zero. 
379: In GTD we interpret the absence of curvature
380:  as a manifestation of the fact that the ideal gas is characterized by the absence of thermodynamic
381: interaction. One of the goals of GTD is to interpret the curvature of the space of equilibrium states as a measure of thermodynamic interaction.
382: This goal has been reached in the case of the ideal gas. Moreover, one can show \cite{quev07,quevaz} that in the case of 
383: the van der Waals gas the curvature tensor of the metric (\ref{gdowns}) does not vanish. Since the van der Waals gas corresponds
384: to a system with non-vanishing thermodynamic interaction, we interpret this result as a further indication that thermodynamic curvature
385: can be used to measure thermodynamic 
386: interaction. In fact, this result has been proved also in the case of more exotic thermodynamic systems like 
387: black holes \cite{quev08grg,aqs08,qs08}.
388: 
389: The metric for the equilibrium space of the ideal gas (\ref{gdownig1}) is invariant with respect to the change of coordinates 
390: $U \rightarrow \lambda U$ and $V \rightarrow \lambda V$, where $\lambda$ is an arbitrary constant. This seems to be a trivial 
391: symmetry related to a rescaling of coordinates. This is true from a mathematical point of view, but physically it implies that 
392: different thermodynamic systems can share the properties of the ideal gas. Indeed, this fact is known from statistical physics. 
393: There are (at least) three different systems which are thermodynamically equivalent: The Maxwell-Boltzmann gas, the Fermi-Dirac gas and
394: the Bose-Einstein gas. Taking  the corresponding partition functions and calculating the relevant variables one
395: can immediately see that they are related by a rescaling of the variables $U$ and $V$ \cite{statistics}. 
396: 
397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
398: \section{Geodesics in the equilibrium space of the ideal gas}
399: \label{sec:geoig}
400: 
401: 
402: Since the curvature of the metric (\ref{gdownig1}) vanishes, there must exist coordinates in which the metric takes the simple Euclidean
403: form. In fact,
404: \be
405: \label{flat}
406: g_S = d\xi^2 + d\eta^2 \ , \quad \xi = \ln U \ , \quad \eta = \ln V \ ,
407: \ee
408: where for simplicity we set the additive constants of integration such that $\xi \geq 0$ and $\eta \geq 0$. This form of the metric is 
409: specially appropriate for the investigation of geodesics. As we mentioned in Section \ref{sec:int}, geodesics of a space represents
410: curves of extremal length. In the case of the equilibrium space ${\cal E}$ we can define the thermodynamic length as \cite{vqs08}: 
411: $L = \int \sqrt{g_S}$. Then, the condition of extremal thermodynamic length, $\delta L =0$, leads to the geodesic equations for the
412: coordinates of the space ${\cal E}$. In the case of the flat metric (\ref{flat}), all the components of the connection vanish and 
413: the geodesic equations become $\ddot \xi =0$ and $\ddot \eta = 0$, where the dot denotes differentiation
414: with respect to the affine parameter $\tau$. 
415: The solutions are found to be  $\xi = \xi_1 \tau+ \xi_0$
416: and   $\eta = \eta_1 \tau + \eta_0$, where $\xi_0,\ \xi_1, \ \eta_0$ and $\eta_1$
417: are constants,  i. e., they represent straight lines.
418: For instance, 
419: consider all geodesics with initial state $\xi_i=0$ and $\eta_i=0$. Then,  
420: on the $\xi\eta-$plane the geodesics 
421: must be contained within the quadrant  determined by  $\xi\geq 0 $ and $\eta \geq 0$, due
422: to our choice of integration constants for $\xi$ and $\eta$ that also fixes the values of
423: the constants $\xi_0,\ \xi_1, \ \eta_0$ and $\eta_1$. Consequently, the geodesics of the ideal gas
424: can be depicted by using the equation $\xi=c_1\eta + c_0$, 
425: with constants $c_0$ and $c_1$. For any arbitrary initial state,  
426: there is always a straight line that connects that state with any arbitrary point on the $\xi\eta-$plane.
427: This means that the entire space of equilibrium states can be covered by geodesics. 
428: This behavior is schematically 
429: illustrated in figure \ref{fig:geo1}.
430: 
431: \begin{figure}
432: \begin{center}
433: \includegraphics[width=7cm]{igfig1.eps}
434: \end{center}
435: \caption{Geodesics in the space of equilibrium states of the ideal gas in logarithmic 
436: coordinates
437: $\xi = \ln U $ and $\eta = \ln V$. The initial state is located on the origin $(0,0)$.  
438: In general, two arbitrary equilibrium states can always be connected by means of a geodesic.}
439: \label{fig:geo1}
440: \end{figure}
441: 
442: However, not all the solutions of the geodesic equations must be physically realistic.  
443: Indeed, there could be straight lines 
444: connecting equilibrium states that are not compatible with the laws of thermodynamics.
445: In particular, one would expect that the second law of thermodynamics imposes strong requirements on the solutions.  
446: In ordinary thermodynamics two equilibrium states are related to each other only if they 
447: can be connected by means of quasi-static process. Then, a geodesic that connects two physically
448: meaningful states can be interpreted as representing a quasi-static process. 
449: Since a geodesic curve is a dense succession of points, we conclude
450: that a quasi-static process can be seen as a dense succession of equilibrium states, a statement
451: which coincides with the definition of quasi-static processes in equilibrium thermodynamics  \cite{callen}.
452: Furthermore, the affine parameter $\tau$ can be used to label all equilibrium states which belong
453: to a geodesic. Since the affine parameter is defined up to a linear transformation, 
454: it should be  possible to choose it  in such a way that it increases as the 
455: entropy of a quasi-static process increases. This opens the possibility of interpreting the
456: affine parameter as a ``time" parameter with a specific direction which coincides with the
457: direction of entropy increase. 
458: 
459: In the special case of the ideal gas, the fundamental equation (\ref{feig}) 
460: in coordinates $\xi$ and $\eta$ represents a straight line 
461: \be
462: \label{feqigs}
463: S = S_0  + N k_{_B} c_{_V} \xi + N k_{_B} \eta \ ,
464: \ee
465: so that the entropy increases as $\xi$ and $\eta$ increases.
466: Consequently, any straight 
467: line pointing outwards of the initial zero point and contained inside 
468: the allowed positive quadrant connect states with increasing entropy. 
469: This behavior is schematically 
470: illustrated in figure \ref{fig:geo2} where the arrows indicate the direction 
471: in which a quasi-static process can take place, i.e., in which the entropy 
472: increases. A quasi-static process connecting
473: states in the opposite direction is not allowed by the second law of thermodynamics. 
474: Consequently, the affine parameter $\tau$ represents a 
475: time parameter and the direction on each geodesic indicates 
476: the ``arrow of time". 
477: \begin{figure}
478: \begin{center}
479: \includegraphics[width=7cm]{igfig2.eps}
480: \end{center}
481: \caption{Geodesics that satisfy the second law of thermodynamics. The initial equilibrium state
482: is located on the origin of coordinates. There is only one geodesic which connects the origin with any 
483: other equilibrium state. The arrows show the direction in which entropy increases, suggesting that they 
484: could be interpreted as the ``arrows of time"'.  
485: }
486: \label{fig:geo2}
487: \end{figure}
488: 
489: 
490: If the initial state is not at the origin of the $\xi\eta-$plane,
491: the second law permits the existence of geodesics for which one of the coordinates,
492: say $\eta$, decreases as long as the other coordinate $\xi$ increases in such a
493: way that the entropy increases or remains constant. 
494: In fact, the region in the $\xi\eta-$plane available from a given initial equilibrium state is 
495: situated within two extreme geodesics which span a maximum angle that can be determined in the following 
496: way. Let the initial state be at the point $(\xi_i,\eta_i)$. According to the second law of thermodynamics and 
497: Eq.(\ref{feqigs}),
498: the state characterized by the coordinate values
499: $(\xi_f, \eta_f)$ can be reached by a geodesic with origin at $(\xi_i,\eta_i)$  
500: if the condition 
501: \be
502: \label{entcond}
503: c_{_V}\Delta \xi  +  \Delta\eta\geq 0 \ ,\quad \Delta \xi = \xi_f -  \xi_i\ ,\quad \Delta \eta = \eta_f -  \eta_i\ ,
504: \ee
505: is satisfied. Consider geodesics for which $\Delta\xi < 0$. Hence, only those geodesics are allowed for which 
506: $\Delta \eta \geq c_{_V}|\Delta\xi|$. The equal sign determines the extreme geodesic with constant entropy which intersects the
507: $\eta-$axis at the point $\xi_f=0$ and $\eta_f = \eta_i + c_{_V}|\Delta\xi| = \eta_i + c_{_V} \xi_i$. This geodesic intersects
508: the  $\eta-$axis at an angle $\alpha$ such that $\tan\alpha = 1/c_{_V}$.  
509: Consider now geodesics with $\Delta\eta < 0$. The allowed
510: geodesics must satisfy $\Delta\xi \geq |\Delta \eta|/c_{_V}$ 
511: and the extreme adiabatic geodesic intersects the $\xi-$axis at 
512: the point with coordinates $\eta_f = 0$ and $\xi_f = \xi_i + |\Delta\eta|/c_{_V} 
513: = \xi_i + \eta_i/c_{_V}$. The adiabatic geodesic 
514: intersects the $\xi-$axis at an angle $\alpha^\prime$ with $\tan\alpha^\prime = c_{_V}$. 
515: Since the intersection angles are 
516: complementary, $\tan\alpha^\prime = 1/\tan\alpha$, we conclude that the angle spanned by  the two adiabatic geodesics (one  
517: with $\Delta\xi <0$ and the second one with  $\Delta\eta <0$) is $\pi$. 
518: 
519: 
520: An alternative derivation of the above geometric construction of adiabatic geodesics consists in considering 
521: the corresponding equation in the form $c_{_V}(\xi_f -  \xi_i)  + \eta_f -  \eta_i = 0$, which can be rewritten 
522: as
523: \be
524: \frac{\xi_f}{\xi_i + \eta_i / c_{_V}} + \frac{\eta_f}{\eta_i + \xi_i c_{_V}} = 1 \ ,
525: \ee
526: and is immediately recognized as the equation of a straight line. This line in the equilibrium space can be occupied only 
527: by states belonging to an adiabatic process. Moreover, since the entropy remains constant along this straight line, the 
528: ``arrow of time" can point in both directions. This is illustrated in figure \ref{fig:geo3}.
529: \begin{figure}
530: \begin{center}
531: \includegraphics[width=7cm]{igfig3.eps}
532: \end{center}
533: \caption [] {Adiabatic geodesics with initial state at $\eta_i=3$, $\xi_i=2$. They represent reversible processes so 
534: that the ``arrow of time" can point in both directions. States in the shadow region are connected to the initial state
535: by geodesics with a negative change in the entropy.
536: }
537: \label{fig:geo3}
538: \end{figure}
539: From the last equation it is then easy to obtain the general relationships
540: \be
541: \tan\alpha = \frac{\xi_i + \eta_i/c_{_V}}{\eta_i + \xi_i c_{_V}}\ ,\quad
542: \tan\alpha^\prime = \frac{\eta_i + \xi_i c_{_V}}{\xi_i + \eta_i/c_{_V}}\ .
543: \ee
544: These formulas are valid for any values of the initial state, except the one situated on the origin of coordinates. 
545: In fact, for the initial 
546: states $(0,\eta_i)$ and $(\xi_i, 0)$, we recover the values of the intersection angles described above. 
547: If the initial state 
548: is on the origin, the entropy condition (\ref{entcond}) is always satisfied since any arbitrary straight line that 
549: starts at the origin is characterized 
550: by $\Delta\xi \geq 0$ and $\Delta\eta \geq 0$ so that the allowed geodesics could occupy the entire positive quadrant as illustrated
551: in figure \ref{fig:geo2}. 
552: However, this result changes drastically if we take into account the third law of thermodynamics which postulates 
553: the impossibility of reaching absolute zero of temperature or, equivalently, the minimum value of the entropy. For the ideal gas with
554: fundamental equation (\ref{feqigs}) the minimum value for the entropy is $S_0$ and corresponds to $\xi=0$ and $\eta=0$. Consequently,
555: the origin of coordinates must be ``removed" from the space of equilibrium states. 
556: 
557: The above results suggest the introduction of the concept of {\it thermodynamic geodesics} as those solutions of the geodesic 
558: equations which satisfy the laws of thermodynamics. In the case of the ideal gas, a thermodynamic geodesic can be represented
559: as a straight line that never crosses
560: the origin of coordinates and possesses a definite direction which  coincides with the direction of entropy increase. 
561: 
562: We see that the laws of thermodynamics imply that the geometric structure of the equilibrium space is as illustrated 
563: in figure \ref{fig:geo4}. For any given initial equilibrium state $(\xi_i,\eta_i)$, there exist two different regions. 
564: The first one is determined by all the states than can be reached from the initial state by means of quasi-static processes, i. e.,
565: by thermodynamic geodesics. This could be called the {\it region of connectivity} of the  initial state. 
566: If we identify $\tau$ as a time parameter, the connectivity region acquires the characteristics of a
567: causally-connected region, resembling concepts of relativistic
568: physics. The second region is composed of all the equilibrium states that cannot be reached from the initial state by thermodynamic
569: geodesics. We call it the {\it region of non-connectivity}. Again, it could be also identified with the non-causally 
570: connected region of spacetime in 
571: relativistic physics. The boundary between the connectivity and non-connectivity regions is occupied 
572: by adiabatic thermodynamic geodesics and this the only place in the equilibrium space where reversible thermodynamic processes can 
573: occur. 
574: 
575: \begin{figure}
576: \begin{center}
577: \includegraphics[width=7cm]{igfig4.eps}
578: \end{center}
579: \caption [] {General structure of the thermodynamic geodesics in the space of equilibrium states of an ideal gas. 
580:  Here we choose a monoatomic gas so that 
581: $c_{_V} = 3/2$. Consequently, $\alpha\approx 33.3^o$ and $\alpha^\prime \approx 56.7^o$.  
582: The shadow region contains all the states that due to the second law of thermodynamics  
583: cannot be reached by thermodynamic geodesics with the fixed initial state. Adiabatic geodesics
584: determine the boundary of the connectivity region where several thermodynamic geodesics are depicted.
585: }
586: \label{fig:geo4}
587: \end{figure}
588: 
589: 
590: It is interesting to mention that the area of the connectivity region is infinite, 
591: whereas the non-connectivity region has a finite area.
592: Indeed, from the above geometric construction (see figure \ref{fig:geo3}) 
593: we see that the non-connectivity region of an arbitrary 
594: initial state $(\xi_i,\eta_i)$ corresponds to a triangle of area
595: \be
596: A_{_{NC}} =  \frac{1}{2c_{_V}}\left(c_{_V}\xi_i + \eta_i\right)^2 \ .
597: \ee
598: Since the heat capacity is related to the number of degrees of freedom per particle, say $l$,  as $c_{_V}=l/2$ 
599: (see, for instance, \cite{statistics}), we see that for ordinary ideal gases with finite $l$ the area of the non-connectivity
600: region remains finite and increases as $l$ grows. If we extrapolate this result to field theories, where the number of degrees 
601: of freedom tends to infinity, the non-connectivity region tends to occupy the entire space of equilibrium states. 
602: This result seems to be interesting in the context of thermodynamics of black holes.  
603: 
604: 
605: 
606: 
607: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%55
608: \section{Geodesics in terms of  thermodynamic variables}
609: \label{sec:geotv}
610: 
611: Since one of the most important properties of GTD is its invariance, we can carry out analysis in arbitrary coordinates. 
612: For the sake of generality, we will present in this section the analysis of the geodesic equations of the space of 
613: equilibrium space of the
614: ideal gas in more conventional thermodynamic coordinates. Consider the flat metric 
615:  (\ref{gdownig1}). According to the formula given in Section \ref{sec:int},
616: the Christoffel symbols for this metric can be written as 
617: \be 
618: \Gamma^U_{\ UU} = -\frac{1}{U} \ , \quad \Gamma^V_{\ VV} = -\frac{1}{V}\ , 
619: \ee
620: and the remaining components vanish identically.
621: Then, the geodesic equations (\ref{geodesics}) become
622: \be
623: \frac{d^2 U }{d\tau^2}  - \frac{1}{U}\left(\frac{dU}{d\tau}\right)^2 = 0 \ , \quad
624: \frac{d^2 V }{d\tau^2}  - \frac{1}{V}\left(\frac{dV}{d\tau}\right)^2 = 0 \ ,
625: \ee
626: where $\tau$ is an affine parameter along the geodesics. These equations can easily be integrated and 
627: yield $U = U_0 e^{\tau/\tau_{_U}},$ and $V = V_0 e^ {\tau/\tau_{_V}}$, respectively,
628: where $U_0$, $V_0$, $\tau_{_U}$ and $\tau_{_V}$ are constants.
629: Then, from these solutions we obtain the relationship
630: \be
631: \label{geoconst}
632: \frac{U^{\tau_{_U}}}{V^{\tau_{_V}}} = const\ ,
633: \ee
634: which allows us to illustrate the form of the geodesics in a simple manner. Some examples
635: are depicted in figure \ref{fig:geo5}.
636: \begin{figure}
637: \begin{center}
638: \includegraphics[width=7cm]{igfig5.eps}
639: \end{center}
640: \caption [] {Geodesics in terms of the thermodynamic variables $U$ and $V$ for
641: different relaxation times $\tau_{_U}$ and $\tau_{_V}$.
642:  Two different types of geodesics are plotted with
643: $\tau_{_U}/\tau_{_V} > 0$ (dotted lines) and $\tau_{_U}/\tau_{_V} < 0$ (solid lines). }
644: \label{fig:geo5}
645: \end{figure}
646: 
647: The relationship (\ref{geoconst}) means that along any geodesic the ratio $U^{\tau_{_U}}/V^{\tau_{_V}}$
648: is a conserved quantity. The explicit values of the constants $\tau_{_U}$ and $\tau_{_V}$
649: depend on the type of process represented by the geodesic. 
650: Consider for instance an adiabatic process for which, according
651: to the fundamental equation for the ideal gas (\ref{feig}), $U^{c_{_V}} V = const$. 
652: Then, a comparison with Eq.(\ref{geoconst}) yields
653: \be
654: \frac{\tau_{_U}}{\tau_{_V}} = - c_{_V} \ .
655: \ee 
656: In the case of a polytropic process with $P V^n = const$, where $n=n(c_{_V})$ is 
657: the polytropic 
658: coefficient, we use the state equations (\ref{steq}) to rewrite (\ref{geoconst}) as
659: $(c_{_V} P)^{\tau_{_U}} V ^{ \tau_{_U}- {\tau_{_V}}} = const$ so that
660: \be
661: n = 1- \frac{\tau_{_V}}{\tau_{_U}}\ .
662: \ee 
663: 
664: The relationship (\ref{geoconst}) is, however, 
665: completely general and the constants ${\tau_{_U}}$ and ${\tau_{_V}}$ do not necessarily 
666: need to be related to the heat capacity $c_{_V}$. This is in agreement with further
667: special cases in which the explicit values of the constants 
668: ${\tau_{_U}}$ and ${\tau_{_U}}$ follow from the analysis of different statistical 
669: ensembles (see, for instance, \cite{statistics}, \S 56). The fact that
670: we obtain the parameters ${\tau_{_U}}$ and ${\tau_{_V}}$ as constants of 
671: integration of the geodesic equations, in which $\tau$ could be interpreted
672: as a ``time" parameter, suggests  the possibility of identifying
673: them as related to the ``relaxation 
674: time" of the thermodynamic variables $U$ and $V$, respectively.
675: 
676: Due to the Legendre invariance of GTD it is possible to perform similar 
677: analysis for any arbitrary thermodynamic potential. In Appendix \ref{sec:app1} we 
678: present all possible thermodynamic potentials and the corresponding metrics
679: for the space of equilibrium states. There, it is shown that in all cases 
680: the solutions to the geodesic equations can be expressed as 
681: $Z^I = Z^I_0 \exp(\tau/\tau_{_I})$, where 
682: $Z^I = (U,V,1/T,P/T)$ and $Z^I_0$ and $\tau_{_I}$ are constants. 
683: Then, we have that
684: \be
685: \frac{(Z^I)^{\tau_{_I}}}{ (Z^J)^{\tau_{_J}}} = const
686: \ee
687: is a universal relationship that holds for all thermodynamic processes of an ideal 
688: gas. It means that for all thermodynamic variables there exist ``relaxation times" 
689: which determine conserved quantities along thermodynamic geodesics. 
690: 
691: It should be 
692: emphasized that we are using the term ``relaxation time" in a very broad sense. 
693: We can interpret $\tau$ as a time parameter  and, consequently, $\tau_{_I}$
694: as a relaxation time, only in the context of quasi-static processes. This 
695: means that we must limit ourselves to the consideration of
696: systems characterized by infinitely slow changes,  measured on a intrinsic time scale,
697: so that after any perturbation the system reaches equilibrium much faster, almost instantaneously, 
698: than its physical parameters vary.
699: 
700: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
701: \section{Conclusions}
702: \label{sec:con}
703: 
704: In this work we present a detailed analysis of the geometry of the ideal gas. The main 
705: property of our analysis is its invariance with respect to Legendre transformations, i. e.,
706: it is independent of the choice of thermodynamic potential, a property which is essential
707: in ordinary thermodynamics. The geometry of the ideal gas turns out to be flat, in 
708: accordance with our intuitive expectation that the absence of thermodynamic interaction
709: would imply absence of curvature. This is an indication that thermodynamic curvature can 
710: be used as a measure of thermodynamic interaction. In fact, this result has been shown to 
711: hold also for more general thermodynamic systems, like the van der Waals gas, and more 
712: exotic systems, like black holes. 
713: 
714: Our analysis of geodesics of the space of equilibrium space shows that they can be represented
715: as straight lines when logarithmic thermodynamic variables are used. We introduce the concept 
716: of thermodynamic geodesics as those solutions of the geodesic equations which satisfy the laws
717: of thermodynamics.  Then, the equilibrium space can be represented as a 
718: Cartesian-like plane where thermodynamic geodesics correspond to quasi-static processes. 
719: The third law of thermodynamics implies that the origin of coordinates must be removed from 
720: the space of equilibrium states. This opens the possibility of interpreting the third law 
721: as a topological property of the equilibrium space. 
722: 
723: For any given initial state, the equilibrium space can be split into two different regions.
724: The connectivity region is occupied by all states which are connected to the initial one by 
725: means of thermodynamic geodesics. On the contrary, the region of non-connectivity corresponds
726: to those equilibrium states than cannot be reached from the initial state by using only 
727: thermodynamic geodesics. In the boundary between these two regions there exist adiabatic 
728: thermodynamic geodesics. It can be shown that this is the only place where adiabatic geodesics
729: can exist so that the boundary determines the only region in the equilibrium space
730: where reversible thermodynamic processes can take place. 
731:    
732: In this work we used the entropy representation. For the sake of completeness, we performed 
733: the same analysis using the energy representation. The main difference in the analysis follows
734: from the fact that the fundamental equation (\ref{feig}) in the energy representation, 
735: $U(S,V) = U_0 \exp(S/N k_{_B}c_{_V})/V^{c_{_V}}$, leads to a non-diagonal term in the metric of the 
736: space  of equilibrium states. This complicates the analysis of the geodesic equations. Nevertheless, 
737: the corresponding metrics with all possible thermodynamic potentials were shown to describe 
738: a flat space so that  in all the cases it is possible to introduce Cartesian-like coordinates 
739: and the investigation of the geodesic equations leads to results equivalent to those obtained in 
740: the entropy representation. This result corroborates one of the most important properties of the
741: formalism of GTD, i. e., the choice of thermodynamic potential or representation does not 
742: affect the results. 
743: 
744: We found that the space of equilibrium states of the ideal gas possesses a very rich geometric 
745: structure which resembles the structure of spacetime in relativistic physics. Since this 
746: structure is the result of applying the laws of thermodynamics in a geometric context, we  
747: expect similar structures in the case of more general thermodynamic systems. For instance, the 
748: metric for the equilibrium space of the van der Waals gas can be obtained immediately 
749: from Eq.(\ref{gdowns}) by introducing the corresponding fundamental equation $S=S(U,V)$. 
750: The resulting metric, however, is no longer flat. This is due to the fact that the van der 
751: Waals gas possesses a certain degree of thermodynamic interaction which, according to the 
752: formalism of GTD, generates a non-vanishing thermodynamic curvature. As a result, the geodesic
753: equations are much more complicated and there are no Cartesian-like coordinates in which the 
754: geodesics could be represented as straight lines. Preliminary calculations indicate that 
755: in this case the geodesic equations must be analyzed by using numerical methods. Due to the 
756: importance of the van der Waals gas, we expect to attack this problem in the future.
757: 
758: In this work we limit ourselves to systems with only two thermodynamic degrees of freedom. 
759: We can increase the number of degrees of freedom, maintaining the flatness of the equilibrium
760: space, by considering, for instance, a multicomponent ideal gas, a paramagnetic ideal gas, etc.
761: In all these case the flatness of the equilibrium space implies that there exist Cartesian-like
762: coordinates. The analysis of thermodynamic geodesics would then require to consider the topological 
763: properties and geometric properties of equilibrium spaces in higher dimensions.
764: 
765: 
766: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
767: \section*{Acknowledgements} 
768: This work was supported in part by Conacyt, Mexico, grants 48601  and 165357. 
769: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
770: \appendix
771: \section{The entropy representation}
772: \label{sec:app1}
773: As mentioned in Section \ref{sec:ig}, in the entropy representation the coordinates
774: of the phase space ${\cal T}$ 
775: are $ (S, U, V, \beta, \vartheta)$, with $\beta =1/T$ and 
776: $\vartheta = P/T$ so that the auxiliary metric in ${\cal T}$ can be written 
777: as
778: \be
779: \label{gups1}
780: G_S = \left(dS -\beta d U - \vartheta dV\right)^2
781: + \Lambda \left[\left(U \beta\right)^{2k+1} dU d\beta  
782: +\left(V\vartheta\right) ^{2k+1} dV d\vartheta \right] \ .
783: \ee
784: As before, the structure of this auxiliary metric is such that 
785: any extensive variable is multiplied by its corresponding intensive variable. As
786: a result the above metric is invariant with respect to the following
787: Legendre transformations,  
788:  $ (S, U, V, \beta, \vartheta) \longrightarrow 
789: (\tilde S, \tilde U, \tilde V, \tilde \beta, \tilde \vartheta)$: 
790: \be
791: \tilde S_1 = S - U \beta \ ,\quad 
792: U = -\tilde \beta \ ,\ \beta = \tilde U \ , 
793: V = \tilde V \ ,\quad  \vartheta = \tilde\vartheta \ ,
794: \ee
795: \be
796: \tilde S_2 = S - V \vartheta \ , \quad
797: U = \tilde U \ , \beta = \tilde \beta \ , \quad
798: V = -\tilde \vartheta\ , \quad \vartheta = \tilde V \ ,
799: \ee
800: \be
801: \tilde S_3 =  S - U \beta - V \vartheta \ , \quad
802: U = -\tilde \beta \ ,\ \beta = \tilde U \ , 
803: V = -\tilde \vartheta\ , \quad \vartheta = \tilde V \ .
804: \ee
805: The thermodynamic potentials $\tilde S_1$, $\tilde S_2$, and $\tilde S_3$ are known 
806: as Massieu functions \cite{callen}. For a given fundamental equation $S=S(U,V)$ 
807: they represent the same subspaces of ${\cal T}$ in different coordinates. 
808: In the case of the ideal gas, the Messieu functions can be derived explicitly 
809: by using the state equations (\ref{steq}) in the form $U\beta = c_{_V} N k_{_B}$ and 
810: $V\vartheta = N k_{_B}$ and the fundamental equation (\ref{feig}). We get (dropping 
811: the tildes) 
812: \be
813: S_1(\beta,V)   = S_{01} - c_{_V}N k_{_B}\ln \beta + N k_{_B} \ln V \ ,
814: \ee
815: \be
816: S_2(U,\vartheta)   = S_{02} + c_{_V}N k_{_B}\ln U -  N k_{_B} \ln \vartheta \ ,
817: \ee
818: \be
819: S_3(\beta,\vartheta)   = S_{03}  - c_{_V}N k_{_B}\ln \beta -  N k_{_B} \ln \vartheta \ ,
820: \ee
821: where $S_{01}$, $S_{02}$, and $S_{03}$ are constants. Then, the metric of the 
822: subspace of equilibrium states ${\cal E}$ in each case corresponds to
823: \be
824: g_1 = \frac{d\beta^2}{\beta^2} + \frac{dV^2}{V^2}\ ,
825: \ee
826: \be
827: g_2= \frac{d U ^2}{U ^2} + \frac{d\vartheta^2}{\vartheta^2}\ ,
828: \ee
829: \be
830: g_3= \frac{d\beta^2}{\beta^2} + \frac{d\vartheta^2}{\vartheta^2}\ ,
831: \ee
832: where we have chosen $\Lambda = -1$ and $k=1$ in the auxiliary metric (\ref{gups1}). Clearly, 
833: all these metrics represent the same flat space of equilibrium states for the ideal
834: gas. The geodesic equations can be solved as explained in Section \ref{sec:ig} and 
835: we obtain $Z^I = Z^I_0 \exp(\tau/\tau_{_I})$, where 
836: $Z^I = (U,V,\beta,\vartheta)$ and $Z^I_0$ and $\tau_{_I}$ are constants. 
837: This is a concrete example of the invariance of the results obtained by using the formalism
838: of GTD.  
839: 
840: 
841: 
842: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
843: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
844: \begin{thebibliography}{99}
845: \bibitem{gibbs} J. Gibbs, {\it The collected works}, Vol. 1, Thermodynamics (Yale 
846: University Press, 1948).
847: 
848: \bibitem{car} C. Charatheodory, {\it Untersuchungen \"uber die Grundlagen der 
849: Thermodynamik},   Math. Ann. {\bf 67}, 355 (1909).  
850: 
851: \bibitem{her} R. Hermann, {\it Geometry, physics and systems} (Marcel 
852: Dekker, New York, 1973). 
853: 
854: \bibitem{wei75} F. Weinhold, {\it Metric Geometry of equilibrium thermodynamics I, II,
855: III, IV, V}, J. Chem. Phys. {\bf 63}, 2479, 2484, 2488, 2496 (1975); {\bf 65}, 558  (1976).
856: 
857: \bibitem{rup79} G. Ruppeiner, {\it Thermodynamics: A Riemannian geometric model},
858: Phys. Rev. A {\bf 20},  1608 (1979).
859: 
860: 
861: \bibitem{rup95} G. Ruppeiner, {\it  Riemannian geometry in thermodynamic fluctuation theory},
862: Rev. Mod. Phys. {\bf 67},  605 (1995); {\bf 68},  313 (1996).
863: 
864: 
865: \bibitem{quev07} H. Quevedo, {\it Geometrothermodynamics}, J. Math. Phys. 
866: {\bf 48}, 013506 (2007). 
867: 
868: \bibitem{nulsal85} J. D. Nulton and P.  Salamon, {\it Geometry of the ideal gas}, Phys. Rev. A {\bf 31},
869: 2520 (1985).
870: 
871: \bibitem{john03} D.A. Johnston, W. Janke, and R. Kenna, {\it Information geometry,
872: one, two, three (and four)}, Acta Phys. Polon. B {\bf 34},  4923 (2003). 
873: 
874: \bibitem{san05c} M. Santoro and Serge Preston, {\it Curvature of the Weinhold metric
875: for thermodynamical systems with 2 degrees of freedom}, (2005); arXiv:math-ph/0505010.
876: 
877: \bibitem{quezar03} H. Quevedo and R. D. Z\'arate, {\it Differential geometry
878: and thermodynamics}, Rev. Mex. F\'\i s. {\bf 49 S2},  125 (2003).
879: 
880: 
881: \bibitem{callen} H. B. Callen, {\it Thermodynamics and an introduction to 
882: thermostatics} (John Wiley \& Sons, Inc., New York, 1985).
883: 
884: \bibitem{arnold} V. I. Arnold, {\it Mathematical methods of classical mechanics}
885: (Springer Verlag, New York, 1980).
886: 
887: \bibitem{vqs08} A. Vazquez, H. Quevedo, and A. Sanchez {\it Thermodynamic systems as 
888: bosonic strings}, (2008); arXiv:hep-th/0805.4819.
889: 
890: \bibitem{burke} W. L. Burke, {\it Applied differential geometry} (Cambridge University
891: Press, Cambridge, UK, 1987).
892: 
893: 
894: \bibitem{quevaz} H. Quevedo and A. V\'azquez, {\it The geometry of thermodynamics}, 
895: AIP Conf. Proc. {\bf 977}, 165  (2008); arXiv:math-ph/0712.0868. 
896: 
897: 
898: \bibitem{quev08grg} H. Quevedo, {\it Geometrothermodynamics of black holes}, 
899: Gen. Rel. Grav. {\bf 40}, 971 (2008).
900: 
901: 
902: \bibitem{aqs08} J. L. \'Alvarez, H. Quevedo, and A. S\'anchez, {\it Unified geometric 
903: description of black hole thermodynamics}, Phys. Rev. D {\bf 77}, 084004 (2008). 
904: 
905: \bibitem{qs08} H. Quevedo and A. S\'anchez, {\it Geometrothermodynamics of asymptotically 
906: de Sitter black holes}, J. High Energy Phys. {\bf 09}, 034 (2008). 
907: 
908: \bibitem{statistics} L. D. Landau and E. M. Lifshitz, {\it Statistical physics} 
909: (Pergamon Press, London, UK, 1980). 
910: 
911: 
912: 
913: 
914: \end{thebibliography}
915: 
916: \end{document}
917: 
918: