1. Introduction
The full thermodynamic description of critical phenomena in the liquid–vapor phase transition of pure substances has remained as a theoretical challenge for a long time [
1]. Substantially, the scaling hypothesis introduced by Widom [
2] proved to be a fundamental step in the understanding of experiments [
3,
4,
5,
6,
7,
8,
9,
10,
11,
12] and numerical simulations [
13,
14,
15,
16] of fluids in the vicinity of the critical point. Of particular interest is Reference [
3], which provides a bibliography of more than 1000 articles, mainly experimental, on critical phenomena between 1950 and 1967. The scaling hypothesis establishes that, if the free energies have a specific functional dependence on their state variables, say Helmholtz free energy in terms of particle density and temperature, the critical exponents are not independent of each other obeying certain equalities [
17,
18,
19]. These exponents characterize the behavior of the thermodynamic properties in the neighborhood of the critical point. Although a consequence of the equalities is that there are two independent exponents only, thermodynamics alone, being an empirical discipline, is unable to predict their numerical values. Indeed, the development of the renormalization group (RG) [
20,
21,
22] led to both a validation of the scaling hypothesis and to a procedure to calculate the exponents in a systematic expansion involving the dimensionality of space. RG in turn is based on certain hypotheses regarding the partition functions of statistical mechanics, mainly scaling invariance close to the critical point. The transcendence of RG, not only in the study of critical phenomena but in many other disciplines, cannot be exaggerated yielding a completely novel approach and understanding of the physics involved. Additionally, one of the major accomplishments of RG concerns the concept of
universality that indicates that the critical exponents are the same not only for all chemically pure fluids but also for all solids showing ferromagnetism, in particular. Perhaps due to these successes, the scaling
hypothesis remained as such from a pure thermodynamic point of view, leaving the impression that thermodynamics alone, with its assumptions based on empirical observations, is truly unable to account for it. From this perspective, the purpose of this article is to show that the scaling hypothesis for the liquid–vapor phase transition can certainly be deduced using thermodynamics only. The present development generalizes the derivation of the scaling hypothesis for the para-ferromagnetic transition presented in Ref. [
23]. As it can be contrasted, the difference between the derivation for a magnetic system, given in such a reference, with the present one for a liquid–vapor transition, is the lack of intrinsic symmetries of the latter, naturally included in the former. Although these results cannot show that the critical exponents have the same values for those two physically dissimilar systems, the whole procedure is based, after all, on the laws of thermodynamics and the phase-equilibrium conditions, which are universal for all substances in nature.
The derivation of scaling starts by analyzing the structure of the entropy per unit of volume
s as a function of the particle density
n and the internal energy per unit of volume
e, namely,
, visualized as a surface on a cartesian
set of axes. The functional dependence of
s on
e and
n is fundamental in the sense that all the equilibrium thermodynamics properties of a pure fluid can be derived from it [
24,
25]. The laws of thermodynamics indicate that
s is a concave single-valued function of
e and
n and that the intensive conjugate variables, temperature
T and chemical potential
, are continuous everywhere. Therefore, the empirical observation of the existence of a liquid–vapor first-order phase transition ending at a critical point, requires that the surface
has a “cut” or void region, such that
e,
n and
s are discontinuous at its edge but
T and
continuous for all pairs of liquid and vapor coexisting states. The edge of such a void region is the
coexistence curve of the transition. The critical point is identified solely as the ending point of the coexistence liquid and vapor states and we make absolutely no additional assumptions about it. As it will be specified, the shape of the surface and the coexistence curve can be quite complicated and, in principle, arbitrary in the
axes, with no prescribed symmetries. However, by changing to a local set of coordinates with the origin at the critical point and along the principal axes of the surface, one can then argue that the coexistence curve is symmetric along the axis tangential to the critical point on the coexistence curve. This assumption is based on the fact that experimental and computer simulated coexistence curves are symmetric very near the critical point [
4,
5,
6,
7,
8,
9,
10,
11,
12,
13,
14,
15,
16]. It is of interest to point out that the analysis of power cycles in the vicinity of the critical point of several substances also supports this assumption [
7]. Not shown here, it is also an exercise to verify that the van der Waals model of a fluid [
24,
25] also shows this symmetry. Concretely, the purpose of this article is to show that these very general considerations on the entropy surface and on the coexistence curve lead to the scaling hypothesis. That is, we show that the functional form of
s on
e and
n, in the vicinity of the critical point, necessarily has the dependence hypothesized by Widom [
2]. A very important but natural consequence of the present analysis is that, while the specific heat at constant volume may or may not diverge at the critical point, the isothermal compressibility necessarily does diverge. We recall that the latter result is equivalent to the appearance of the unbounded density fluctuations and of the losing of all length scales at the critical point, which are the essence of RG [
20,
21,
22]. It is thus very exciting to find out that thermodynamics predicts this divergent behavior without appealing to the molecular structure of the fluid. A geometrically equivalent way to express these critical divergences is that the existence of a curve of coexistence of liquid–vapor phases, bounding a void region on an otherwise concave function, implies the vanishing of the gaussian curvature of the surface at the critical point; that is, the surface forcibly becomes locally flat at such a point.
In
Section 2 we present a brief summary of the general thermodynamic properties of the function
.
Section 3 is devoted to the isometrical transformation from the natural axes
to an appropriate local set of coordinates at the critical point, in which one of the axes is the normal to the surface, other is the tangent to the coexistence curve at the critical point, with the third one being orthogonal to the previous ones. In
Section 4 we analyze the strong requirements that the coexistence curve imposes on the local entropy function and its derivatives, and show that these conditions straightforwardly imply scaling of the entropy function. We also discuss the general properties of the obtained scaling functions of the entropy.
Section 5 is dedicated to the derivation of the usual critical exponents for the behavior of the density and chemical potential in terms of the temperature, as well as for the specific heat and constant volume and the isothermal compressibility, near the critical point. We conclude with some final remarks that we consider to be relevant. Details of some lengthy calculations and a generalization of the derivation of the scaling forms are given in two appendices.
2. Thermodynamic Conditions for the Liquid–Vapor Phase Transition in a Pure Fluid
Let us consider the entropy
of an “arbitrary” chemically pure fluid.
s,
e and
n are the entropy, energy and number of particles per unit of volume. By the laws of thermodynamics,
is a single valued, concave function of
, such that [
24,
25],
with all the variables in dimensionless units (say, entropy in units of Boltzmann constant and energy and volume with units of two characteristic parameters of intermolecular potentials, see the Nomenclature at the end of the article).
and
, with
T the temperature and
the chemical potential. By the third law
, which indicates that for
constant
s is a concave, monotonic increasing function of
e. Although there is no thermodynamic restriction on
, for states near the liquid–vapor transition
[
26] and, as a consequence,
s is also a concave, monotonic increasing function of
n, for
constant, see
Figure 1.
The pressure
p of the fluid is given by the Euler relationship, [
24,
25]
Thermodynamic equilibrium requires that
,
and
p are
continuous functions of
. In addition, the second law guarantees that the principal curvatures of the surface
are finite everywhere, except at isolated points such as the critical one. The fact that
is concave everywhere yields the stability conditions on the specific heat at constant volume and number of particles
and on the isothermal compressibility
[
24,
25],
and
Now, we consider a fluid that shows a liquid–vapor phase transition ending in a thermodynamic state known as the
critical point, see
Figure 1. In such a phase transition, except at the critical point, there exists a continuum of pairs of thermodynamic states in equilibrium, with their energy
e, particle number
n, and entropy
s densities being discontinuous. This physical situation requires that the entropy function
, considered a surface in a cartesian set of axes
, shows a “cut”, or void, that accounts for the mentioned discontinuities. The edge of such a void region is the
coexistence curve, as shown in
Figure 2. For values of
“inside” the void
s is not defined. The curve has a special point, identified as the critical one
, such that for a given pair of the mentioned states, one is the liquid phase with values
in one side of the critical point, and the other is the gas phase with
in the opposite side. These states are said to be in coexistence if their temperature
and chemical potential
(and so pressure
p) have the same values. As the critical point is approached the two coexisting states coalesce into such a special point. Here, we make the
unique assumption of this discussion, based entirely on experimental and numerical simulations data [
3,
4,
5,
6,
13,
14,
15,
16]: very near the critical point, including the coexistence curve, the surface is symmetric with respect to the plane perpendicular to the tangent at the critical point, as it will be explicitly specified below. This very important empirical observation will lead to the scaling form of
and to the well-known critical properties, namely, the necessary divergence of the isothermal compressibility and the possible divergence of the specific heat at constant volumen. A very important consideration is that the entropy surface is an analytic function of
, except at the critical point where it can be non-analytic.
3. An Isometric Transformation to the Critical Point
As mentioned above, the entropy function
can be considered as a surface
in the right-hand set of axes
. Now, for our purposes below, we make a transformation to a cartesian set of axes
located at the critical point
, as shown in
Figure 3. The
axis is defined by the normal unit vector at the critical point,
where
and
are the (inverse) temperature and chemical potential evaluated at the critical point. In the following, all quantities with subindex “
c” are meant to be evaluated at the critical point. The
axis in turn is defined by the unit vector tangent to the coexistence curve, at the critical point,
, and the
axis is then identified by the vector
, pointing towards the region where the surface
is defined. The relationship between the coordinates
and the local set
is given by,
where
with
,
and
.
In the new set of coordinates
, the entropy surface can be expressed in terms of the function
, see third relationship in Equation (
6), and thus is related to the entropy by,
where
and
are given by the first two equations of Equation (
6), with
.
With the given transformation, the relations between
,
and the derivatives of
z with respect to
x and
y are,
The following is a general result: from Equation (
6) one finds that the derivatives of
x and
y with respect to
e depend on
only, while the derivatives of
x and
y with respect to
n, in turn, depend on
only. Therefore, it is a simple exercise to verify that for two coexisting states, the fact that
and
have the same value implies that the derivatives
and
must also have the same value for those two coexisting states.
Our interest is the description of the entropy in the vicinity of the critical point, namely, for
and
. Following our physical assumption that the coexistence curve is symmetric with respect to the tangent
at the critical point, we can assert that in this set of coordinates the coexistence curve is given by a relationship between
x and
y, namely
, such that the coexistence curve can be written parametrically through the vector,
with
a symmetric function of
y,
with
, a coefficient characteristic of the given fluid, and
an exponent that, one expects, should be universal, but that at the level of thermodynamics it is determined by the function
. It is physically reasonable to assume that
, since the coexistence curve should have a curvature either finite or zero at the critical point. Since
is not limited to be an integer, the coexistence curve can be non-analytic at the critical point. The coexistence curve is represented by the solid (red) line in
Figure 3. From now on, we shall use the notation
for any
, to avoid cumbersome expressions.
The assumed form of the coexistence curve, Equation (
11), implies that, in the vicinity of the critical point, the thermodynamic states
and
coexist. Hence, the temperature
and the chemical potential
have the same value at those states, that is,
and
. Now, we make the most important assumption of the present work: very near the critical point, that is, at a leading order, the function
is symmetric on
y,
. As mentioned in the Introduction, one can show that the van der Waals model indeed satisfies this requirement. From this assumption follows a transcendental result. First, note that the even symmetry of
z on
y implies that the derivative of
z with respect to
y is odd. However, as stated above, both derivatives of
z with respect to
x and
y must be equal at coexistence states. Therefore, since the derivative of
z with respect to
y at coexistence must be both odd and even, this can only be true if it vanishes at all points in the coexistence curve, that is,
As we now show, this condition on the shape of the surface is so strong that it implies that must obey scaling.
4. The Scaling Form of the Entropy
A very important condition is that the entropy surface is analytic everywhere, except perhaps at the critical point. Therefore, we can make an
x-power expansion of
around
, for an arbitrary value of
, near the critical point. This yields,
where the functions
need not be analytic at
. Let us now take the derivative of
z with respect to
y,
where the prime means differentiation with respect to the argument. As we have just argued above, see Equation (
12), this derivative must be zero at the coexistence curve, that is, for
,
We note that this condition imposes a very strong restriction on the functions
since the equality must be true for a continuum of values of
. Although we are considering this to hold near the critical point only, we can write a quite general expression that satisfies the above requirement. That is, a general solution to Equation (
15) is that
is a power law expansion, not necessarily analytic:
where the exponents
,
,
, and so on, are not integers in general. With this proposal we find that for expression (
15) to hold, these exponents must satisfy,
and so on, with
, etcetera, with no loss of generality. To verify it, substitute the above into Equation (
15),
from which one concludes that to satisfy the equality for all values of
y, each sum must vanish separately. In particular, and our interest here, it must be true that,
Hence, substituting Equations (
16) and (
17) into Equation (
13), we can write,
The above form shows a general functional dependence that obeys scaling. However, since we have limited ourselves to assumptions and approximations very near the critical point, we keep the lowest order of
only. This yields the desired scaling expression for the surface function
in terms of two unknown exponentes
and
,
where in the last line we have defined the scaling function
, with
, which by construction is an analytic function of its argument. This demonstrates that the entropy can be written in a scaling form in terms of the variables
x and
y [
2,
19,
21]. In the following section we analyze the predictions of this finding and in
Appendix A we show that the series expansion given in Equation (
13) can be made even more general, leading too to the scaling form in Equation (
21). Before writing the full expression of the entropy, we first address how to deal with values of
and
; this is discussed in many reviews, such as in Ref. [
19].
Because
is a maximum at the origin, then
. For
,
, where
. While the scaling function is valid everywhere
on the entropy surface, it can be explicitly evaluated for
but
only, as expressed above. However, for
and
,
z must be a function of
x only. Therefore, asymptotically, the scaling function must behave as,
with
, such that, see Equation (
21),
, a function of
x only. This indicates that, in general [
19], as long as
, we can “invert” the function
by identifying a new function as,
and since
can be non analytic at criticality, this scaling function should also be an infinite series,
Conversely and for consistency, asymptotically it must be true that,
Therefore, one can write
in terms of this complementary scaling function,
This form is particularly useful to calculate properties of the surface in the limit
for
, namely, at the states on the dotted (blue) line in
Figure 3, that we can now identify as the line of “symmetry breaking” [
21,
22]. This is because we can now show that the derivative of
z with respect to
y, at
and for any
, vanishes:
That is, for
and
there is only one phase (a so-called supercritical fluid) with the derivative
being zero; then for
but at the coexistence curve, there are two phases with
remaining zero. The analogy is with the para-ferromagnetic case where the line of symmetry breaking is when the magnetic field vanishes [
21,
23]; that is, in the local frame
the derivative of
z with respect to
y is the analog of the magnetic field. We return to this relation in the final section.
With the previous identification of the scaling function we can write explicit forms of the entropy in the vicinity of the critical point,
valid for all values of
and
, or its alternative form
valid for
and all values of
y. Consistently,
x and
y, given by Equation (
6) in terms of
and
, should also be expanded up to the appropriate order to yield the leading order of the thermodynamic properties in question.
To conclude this section, we can observe the general properties of the scaling functions. Since
is a maximum at
, then
and
for all values of their arguments. For positive values of
,
. However, at coexistence
must be negative; as it will be seen below, this is due to the fact that the (inverse) temperature
at coexistence states is greater than the critical (inverse) temperature
. As a consequence, at a certain negative value of its argument
where it changes sign. As we will show below,
occurs at the critical isotherm. See
Figure 4 for a sketch of
.
6. Final Remarks
The derivation of the scaling form of the entropy here presented partially follows the study of Ref. [
23] for a ferromagnetic system. However, due to the much more complicated structure of the entropy function of a fluid near its critical point, the study here includes the magnetic one as an special case. In that situation, the entropy per unit volume
s is a function of the energy per unit volume
e and the magnetization per unit volume
m, that is,
. By the laws of thermodynamics
s is a concave function of those variables. In this case,
where
H is the magnetic field. By physical reasons of symmetry,
s is an even function of
m and, therefore,
is an odd function of
m. The critical point is at
,
and
. Thus, the symmetry breaking occurs for
where, with
, one finds
. Hence, since
s is even in
m and the critical value of
H is zero, the identification
in our discussion follows right away. This suggests that the isometric transformation Equation (
6) is actually unnecessary, yielding
. Therefore, one immediately finds that the entropy can be written near the critical point as,
with
the coexistence curve and
a scaling function [
23]. It is worth recalling that when analogies between magnetic and fluid systems are considered, see, for example, Refs. [
18,
21], the magnetization
m is the analog of density
and it appears natural to identify the magnetic field
H as the analog of the chemical potential
(or the pressure) since these are the thermodynamic conjugate variables of the former. However, the present study shows that this is not the case. That is, the correct analogy of the magnetic symmetry-breaking states with
, are in a fluid those for which the derivative
vanishes, given by Equation (
27), as discussed in
Section 2. That is, very near but above the critical point, Equation (
27) leads to the symmetry-breaking line in the fluid, defined by
while below the critical point, the line is the coexistence curve
, or the expression given by Equation (
32) in terms of density and temperature. As a matter of fact, this was already implicitly discussed by Widom in Ref. [
2].
Although not shown here, an illustrative and pedagogical exercise is the analysis of the mean-field van der Waals fluid [
24], in the light of the present study. In this case, one can find the explicit isometric transformation given by Equation (
6) and work out and verify all the predictions here discussed. As one can expect, van der Waals and ferromagnetism Landau mean field models correspond to
and
. Details will be given elsewhere.
We find highly interesting to point out that while the divergence of the compressibility at the critical point is a physical result with profound and transcendental consequences, notably the foundation of the renormalization group description of the critical point, it appears here to follow as a geometric constraint of the entropy surface at the critical point. That is, the condition
, which ultimately follows from the second law of thermodynamics, indicates “simply” that the gaussian curvature of the entropy surface, being positive and finite at any stable thermodynamic state, becomes zero at the critical point. This implies that, at least, one of the eigenvalues of the surface curvature tensor vanishes at the critical point, if not both. This general result can be obtained from pure geometrical arguments without appealing to scaling. In other words, both scaling and the flatness of the entropy surface at the critical point, follow from the equilibrium conditions at coexistence and from the fact that there is an ending point to such a coexistence. To provide evidence for this, first we recall that the equilibrium conditions of having the same temperature and chemical potential (and thus pressure) for a given pair of coexistence states, is equivalent to assert that while the entropy is discontinuous at the first order phase transition coexistence curve, the surface normal vector
is the same at both coexisting states. Therefore, since such a pair of normal vectors must coalesce to the critical normal
staying parallel between them, it is not difficult to show that this implies that the surface necessarily becomes flat at the critical point, at least along one of its principal directions. The case where only one of the curvature eigenvalues is zero corresponds to
and
, the mentioned mean-field case to which Landau and van der Waals theories belong [
19,
21]. For
both eigenvalues are explicitly zero. There is a borderline case, not considered here but amply discussed in the seminal paper by Widom [
2], in which
, such as the two-dimensional Ising model [
18,
29], where
logarithmically at the critical point, and both eigenvalues also vanish. One can further verify that the critical exponents for this case correspond to
and
.
To conclude we would like to speculate about going further with a pure thermodynamics search for the elucidation of the actual values of the critical exponents. As we have discussed here, from the laws of thermodynamics and the reasonable assumption of a local even symmetry in
y in the vicinity of the critical point, scaling follows as a consequence of the phase coexistence requirement. The most that we can conclude so far is that there exist only two seemingly independent “unknown” exponents
and
. This imposes the condition that the critical exponents are determined by two seemingly independent exponents
and
, which in turn may be characteristic of the given fluid, being encoded in the entropy function
. However, we do know that there exists universality, namely, that the exponents are the same for all pure substances that show a liquid–vapor critical point. Hence, it should be noticed that from a pure thermodynamic perspective, such as the present one and all the others, see Refs. [
2,
21,
22], it appears that universality cannot be accounted for. We suggest that this may not be case. On the one hand, the observation that there are only two independent exponents is certainly in agreement with RG, the theory that provides a systematic procedure to calculate those two exponents [
21]. On the other hand, the present study shows that scaling has its root in the geometric constraint that the coexistence curve imposes on the necessary discontinuity of the entropy surface. Hence, a question arises as to whether there could be additional constraints that would force the exponents
and
, first of all be universal and, second, not independent of each other. That is, that the mere facts that the surface must be concave and that the curve must be of coexistence, namely, that derivatives of the surface must be continuous across such a curve, may indicate that only a particular set of surfaces can satisfy those constraints. As a possible inkling on this, we highlight here once more that the coexistence condition that leads to scaling is the requirement that the derivative of the function
z with respect to
y vanishes at the coexistence curve,
. This yields the interesting condition, see Equation (
21),
This is an unexpected result in the context of the universality of the exponents
and
, since it is usually believed that their values should be independent of the precise functional form of the scaling function
. The above expression apparently indicates that
and
are not independent, although it may just be an identity showing the asymptotic behavior of
near the coexistence curve, as illustrated in
Figure 4. We believe it is worthwhile to explore possible forms of the scaling function in terms of concave surfaces with discontinuities, to find out whether it is indeed an identity or it is a gate to find an additional thermodynamic critical exponents relationship. The truthfulness of this would imply a very deep consequence, for then only one exponent would be necessary to find out, the rest of them following by scaling.