7 ▪ Phase transitions and critical phenomena


CHAPTER 7


Phase transitions and critical phenomena


PHASE, as the term is used in thermodynamics, refers to a spatially uniform equilibrium system.1 A body of given chemical composition can exist in a number of phases. H2O, for example, can exist in the familiar liquid and vapor phases, as well as several forms of ice having different crystal structures. Substances undergo phase transitions, changes in phase that occur upon variations of state variables. Phases can coexist in physical contact (such as an ice-water mixture). Figure 7.1 is a generic phase diagram, the values of T,P for which phases exist and coexist along coexistence curves. Note the triple point, a unique combination of T,P at which three phases coexist.2 At the critical point, the liquid-vapor coexistence curve ends at Tc,Pc (critical temperature and pressure), where the distinction between liquid and gas disappears.3 For T>Tc, a gas cannot be liquefied regardless of pressure. In the vicinity of the critical point (the critical region), properties of materials undergo dramatic changes as the distinction between liquid and gas disappears. Lots of interesting physics occurs near the critical point—critical phenomena—basically the remainder of this book.4

Figure 7.1

Figure 7.1Phase diagram in the PT plane. Thick lines are coexistence curves.


 



7.1.1 Equilibrium conditions


We start by asking whether there is a limit to the number of phases that can coexist. An elegant answer is provided by the Gibbs phase rule, Eq. (7.13). The chemical potential of substances in coexisting phases has the same value in each of the phases in which coexistence occurs.5 Consider two phases of a substance, I and II. Because matter and energy can be exchanged between phases in physical contact, equilibrium is achieved when T and P are the same in both phases, and when the chemical potentials are equal, μI=μII (see Section 1.12). We know from the Gibbs-Duhem equation,6 (P1.1), that μ=μ(T,P), and thus chemical potential can be visualized as a surface μ=μ(T,P) (see Fig. 7.2). Two phases of the same substance coexist when


μI(T,P)=μII(T,P).


(7.1)


The intersection of the two surfaces defines the locus of points P=P(T) for which Eq. (7.1) is satisfied—the coexistence curve (see Fig. 7.2). Three coexisting phases (I,II,III) would require the equality of three chemical potential functions,


μI(T,P)=μII(T,P)=μIII(T,P).


(7.2)

Figure 7.2

Figure 7.2Coexistence curve defined by the intersection of chemical potential surfaces associated with phases I and II of a single substance.


Equation (7.2) implies two equations in two unknowns and thus three phases can coexist at a unique combination of T and P, the triple point. By this reasoning, it would not be possible for four phases of a single substance to coexist (which would require three equations in two unknowns). Coexistence of four phases of the same substance is not known to occur.


Multicomponent phases have more than one chemical species. Let μjγ denote the chemical potential of species j in the γ phase (we use Roman letters to label species and Greek letters to label phases). Assume k chemical species, 1jk,, and π phases, 1γπ,, where π is an integer. The first law for multiphase, multicomponent systems is the generalization of Eq. (1.21):


 


dU=TdSPdV+γ=1πj=1kμjγdNjγ.


(7.3)


Extensivity implies the scaling property U(λS,λV,λNjγ)=λU(S,V,Njγ). By Euler’s theorem,


U=SUSV,Njγ+VUVS,Njγ+γ=1πj=1kNjγUNjγS,V=TSPV+γ=1πj=1kμjγNjγ,


(7.4)


where the derivatives follow from Eq. (7.3) and Njγ¯ indicates to hold fixed particle numbers except Njγ. Equation (7.4) implies


G=γ=1πj=1kμjγNjγ,


(7.5)


the generalization of Eq. (1.54). Taking the differential of Eq. (7.4) and using Eq. (7.3), we have the multicomponent, multiphase generalization of the Gibbs-Duhem equation, (P1.1),


γ=1πj=1kNjγdμjγ=SdT+VdP.


(7.6)


By taking the differential of G in Eq. (7.5), and making use of Eq. (7.6),


dG=SdT+VdP+γ=1πj=1kμjγdNjγ,


(7.7)


and thus we have the alternate definition of chemical potential,7 μjγ=G/NjγT,P,Njγ¯, the energy to add a particle of type j in phase γ holding fixed T, P, and the other particle numbers.


The Gibbs energy is a minimum in equilibrium.8 From Eq. (7.7),


dGT,P=γ=1πj=1kμjγdNjγT,P=0.


(7.8)


If the particle numbers Njγ could be independently varied, one would conclude from Eq. (7.8) that μjγ=0. But particle numbers are not independent. The number of particles of each species spread among the phases is a constant,9 γNjγ= constant, and thus there are k equations of constraint


γ=1πdNjγ=0.  j=1,,k


(7.9)


 


Constraints are handled through the method of Lagrange multipliers10 λj, which when multiplied by Eq. (7.9) and added to Eq. (7.8) leads to


j=1kγ=1πμjγ+λjdNjγ=0.


(7.10)


We can now treat the particle numbers as unconstrained, so that Eq. (7.10) implies


μjγ=λj.


(7.11)


The chemical potential of each species is independent of phase. Equation (7.11) is equivalent to


μj1=μj2==μjπ.    j=1,,k


(7.12)


There are k(π1) equations of equilibrium for k chemical components in π phases.


7.1.2 The Gibbs phase rule


How many independent state variables can exist in a multicomponent, multiphase system? In each phase there are Nγj=1kNjγ particles, and thus there are k1 independent concentrations cjγNjγ/Nγ, where j=1kcjγ=1. Among π phases there are π(k1) independent concentrations. Including P and T, there are 2+π(k1) independent intensive variables.


There are k(π1) equations of equilibrium, Eq. (7.12). The variance of the system is the difference between the number of independent variables and the number of equations of equilibrium,


f2+π(k1)k(π1)=2+kπ.


(7.13)


Equation (7.13) is the Gibbs phase rule.[11, p96] It specifies the number of intensive variables that can be independently varied without disturbing the number of coexisting phases ( f0).


k=1,π=1f=2: a single substance in one phase. Two intensive variables can be independently varied; T and P in a gas.


k=2,π=1f=3: two substances in a single phase, as in a mixture of gases. We can independently vary T, P, and one mole fraction.


k=1,π=2f=1: a single substance in two phases; a single intensive variable such as the density can be varied without disrupting phase coexistence.


k=1,π=3f=0: a single substance in three phases; we cannot vary the conditions under which three phases coexist in equilibrium. Unique values of T and P define a triple point.


One should appreciate the generality of the phase rule, which doesn’t depend on the type of chemical components, only that the Gibbs energy is a minimum in equilibrium.


7.1.3 The Clausius-Clapeyron equation


The latent heat, L, is the heat released or absorbed during a phase change, and is measured either as a molar quantity (per mole) or as a specific quantity (per mass). One has the latent heat of vaporization (boiling), fusion (melting), and sublimation.11 Latent heats are also called the enthalpy of vaporization (or fusion or sublimation).12 At a given T,


L(T)=hvhl=T sv(T,P(T))sl(T,P(T)),


(7.14)


 


where v and l refer to vapor and liquid, lower-case quantities such as sS/n indicate molar values, and P(T) describes the coexistence curve.13 The difference in molar entropy between phases is denoted Δssvsl; likewise with Δh. Equation (7.14) is written compactly as L=Δh=TΔs.


As we move along a coexistence curve, T and P vary in such a way as to maintain the equality μI(T,P)=μII(T,P). Variations in T and P induce changes δμ in the chemical potential, and thus along the coexistence curve, δμI=δμII. For a single substance, dμ=sdT+vdP (Gibbs-Duhem equation). At a coexistence curve, sIdT+vIdP=sIIdT+vIIdP, implying


dPdTcoexist=sIsIIvIvII=ΔsΔv=ΔhTΔv=LTΔv.


(7.15)


Equation (7.15) is the Clausius-Clapeyron equation; it tells us the local slope of the coexistence curve. If one had enough data for L as a function of T, P and the volume change Δv, Eq. (7.15) could be integrated to obtain the coexistence curve in a PT diagram.



Phase coexistence at a given value of (T,P) requires the equality μI(T,P)=μII(T,P), but that says nothing about the continuity of derivatives of μ at coexistence curves. For a single substance, μ=G/N, and thus from Eq. (1.14) (or the Gibbs-Duhem equation (P1.1)),


VN=μPT,N    SN=μTP,N.


(7.16)


For second derivatives, using Eqs. (1.33) and (P1.6), together with Eq. (7.16),14


βTN=1V(V/N)PT,N=1V2μP2T,NCPN=T(S/N)TP,N=T2μT2P,N.


(7.17)


The behavior of these derivatives at coexistence curves allows a way to classify phase transitions.


• If the derivatives in Eq. (7.16) are discontinuous at the coexistence curve (i.e., μ(T,P) has a kink at the coexistence curve), the transition is called first order and the specific volume and entropy are not the same between phases, Δv0, Δs0. The Clausius-Clapeyron equation is given in terms of the change in specific volume Δv and entropy, Δs=L/T. If a latent heat is involved, there are discontinuities in specific volume and entropy, and the transition is first order.


• If the derivatives in Eq. (7.16) are continuous at the coexistence curve, but higher-order derivatives are discontinuous, the transition is called continuous. For historical reasons, continuous transitions are also referred to as second-order phase transitions.15 At continuous phase transitions, Δv=0 and Δs=0 —there is no latent heat. Entropy is continuous, but its first derivative, the heat capacity, is discontinuous. Whether or not there is a latent heat seems to be the best way of distinguishing phase transitions.


 



The van der Waals equation of state, (6.32), is a cubic equation in the specific volume, V/N:


VN3b+kTPVN2+aPVNabP=0.


(7.18)


Cubic equations have three types of solutions: a real root together with a pair of complex conjugate roots, all roots real with at least two equal, or three real roots.[50, p17] The equation of state is therefore not necessarily single-valued;16 it could predict several values of V/N for given values of P and T. For large V/N (how large?—see below), we can write Eq. (7.18) in the form


(VN)2[VN(b+kTP)+aN2PV2(VNb)]=0.


(7.19)


Large V/N is the regime where the volume per particle is large compared with microscopic volumes, V/Na/(kT),b (the van der Waals parameter a has the dimension of energy-volume). In this limit, the nontrivial solution17 of Eq. (7.19) is the ideal gas law, PV=NkT. For small V/N, we can guess a solution of Eq. (7.18) as V/Nb (the smallest volume in the problem), valid at low temperature; see Exercise 7.3. (We can’t have V/N=b, which implies infinite pressure.) For V/Nb, densities approach the close-packing density. We have no reason to expect the van der Waals model to be valid at such densities, yet it indicates a condensed phase, an incompressible form of matter with V/Nconstant, independent of P. At low temperature, kTPb, Eq. (7.18) has one real root, with a pair of complex conjugate roots (see Exercise 7.4).


7.3.1 The van der Waals critical point


Thus, we pass from the low temperature, high density system (low specific volume), in which Eq. (7.18) has one real root, to the high temperature, low density system (large specific volume) which has one nontrivial root (ideal gas law). We shouldn’t be surprised if in between there is a regime for which Eq. (7.18) has multiple roots (which hopefully can be interpreted physically). Detailed studies show there is a unique set of parameters Pc, Tc, and (V/N)c (which as the notation suggests are the pressure, temperature, and specific volume of a critical point18), at which Eq. (7.18) has a triple root. For (V/N)cvc a triple root of Eq. (7.18), the cubic equation


vvc3=v33vcv2+3vc2vvc3=0


(7.20)


must be equivalent to Eq. (7.18) evaluated at Pc and Tc. Comparing coefficients of identical powers of v in Eqs. (7.18) and (7.20), we infer the correspondences


3vc=b+kTcPc    3vc2=aPc    vc3=abPc.



These equations imply the values of the critical parameters (show this):


vc=3b    Pc=a27b2    kTc=8a27b.


(7.21)


The relations in Eq. (7.21) predict Pc and Tc reasonably well, but overestimate vc (Exercise 7.5). Equation (7.21) also predicts Pcvc/(kTc)=38 (show this), a universal number, the same for all gases. In actuality, Pcvc/(kTc)0.2 –0.3 for real gases.19 Should we be discouraged by this disagreement? Not at all! The van der Waals model is the simplest model of interacting gases—if we want better agreement with experiment, we have to include more relevant physics in the model.


 


7.3.2 The law of corresponding states and the Maxwell equal area rule


Express P, T, and v in units of the critical parameters. Let


P¯P/Pc    T¯T/Tc    v¯v/vc.


(7.22)


The dimensionless quantities P¯,T¯,v¯ are known as the reduced state variables of the system. With P=PcP¯, T=TcT¯, and v=vcv¯, the van der Waals equation of state can be written (show this)


(P¯+3v¯2)(3v¯1)=8T¯.


(7.23)


Equation (7.23) is a remarkable development: It’s independent of material-specific parameters, indicating that when P, T, v are scaled in units of Pc, Tc, vc (which are material specific), the equation of state is the same. Systems having their own values of (a,b), yet the same values of P¯, T¯, v¯, are said to be in corresponding states. Equation (7.23) is the law of corresponding states: Systems in corresponding states behave the same. Thus, argon at T=300 K, for which Tc=151 K, will behave the same as ethane (C2H6) at T=600 K, for which Tc=305 K (both have T¯2).20 Figure 7.3 is a compelling illustration of the law of corresponding states. For ρl (ρg) the density of liquid (vapor) in equilibrium at T<Tc, by the law of corresponding states we expect ρl/ρc and ρg/ρc to be universal functions of T/Tc, where ρc denotes the critical density. Figure 7.3 shows data for eight substances (of varying complexity21 from noble-gas atoms Ne, Ar, Kr, and Xe, to diatomic molecules N2, O2, and CO, to methane, CH4), which, when plotted in terms of reduced variables, ostensibly fall on the same curve![83] The solid line is a fit to the data points of Ar (solid circles) assuming the critical exponent β=13 (see Section 7.4). Figure 7.3 shows the coexistence region—values of temperature and density for which phase coexistence occurs.22

Figure 7.3

Figure 7.3The law of corresponding states. Coexistence regions are the same when plotted in terms of reduced state variables. Reproduced from E.A. Guggenheim, J. Chem. Phys. 13, p. 253 (1945), with the permission of AIP Publishing.


Because the equation of state is predicted to be the same for all systems (when expressed in reduced variables), any thermodynamic properties derived from it are also predicted to be the same. Universality is the term used to indicate that the behavior of systems near critical points is largely independent of the details of the system.23 Universality has emerged as a key aspect of modern theories of critical phenomena, as we’ll see in Section 8.4.24


Figure 7.4 shows the isotherms calculated from Eq. (7.23). For TTc, there is a unique specific volume for every pressure—the equation of state is single valued.25 For T<Tc, however, there are three possible volumes associated with a given pressure, such as the values of v/vc shown at A,B,C for T=0.9Tc. How do we interpret these multiple solutions?

Figure 7.4

Figure 7.4Isotherms of the van der Waals equation of state. Points A,B, and C are three possible volumes associated with the same pressure (and temperature).


 


Figure 7.5 shows an expanded view of an isotherm for T<Tc. We note that (P/v)T is positive along the segment bcd, which is unphysical. An increase in pressure upon an isothermal expansion would imply a violation of the second law of thermodynamics. The root c in Fig. 7.5 (or B in Fig. 7.4) is unphysical and can be discarded.

Figure 7.5

Figure 7.5Subcritical isotherm of the van der Waals equation of state ( T=0.85Tc), with its unphysical segment bcd. The point c (root of the equation of state) is unphysical and can be discarded.


What to make of the other two roots? Figure 7.6 indicates what we expect phenomenologically and hence what we want from a model of phase transitions. As a gas is compressed, pressure rises.26 When (for T<Tc) the gas has been compressed to point B, condensation first occurs (coalescence of gas-phase atoms into liquid-phase clusters). At B, the specific volume vg represents the maximum density a gas can have at T<Tc. Upon further isothermal compression, the pressure remains constant along the horizontal segment AB (at the saturation pressure, Psat(T)) as more gas-phase atoms are absorbed into the liquid phase.27 At point A, when all gas-phase atoms have condensed into the liquid, the specific volume v=vl is the minimum density a liquid can have at T<Tc. Further compression at this point results in a rapid rise in pressure.

Figure 7.6

Figure 7.6Liquid-gas phase diagram in the Pv plane. Dotted line is the prediction of the van der Waals model. The horizontal segment AB of the isotherm indicates liquid-gas coexistence.


 


At a point in the coexistence region, such as point D in Fig. 7.6, there are Nl atoms in the liquid phase and Ng in the vapor phase, where N=Nl+Ng is the total number of atoms. Let xlNl/N and xgNg/N denote the mole fractions of the amount of material in the liquid and gas phases. Clearly xl+xg=1. The volume occupied by the gas (liquid) phase at this temperature is Vg=Ngvg ( Vl=Nlvl), implying V=Ngvg+Nlvl. The specific volume vd associated with point D is:28


vd=VN=Ngvg+NlvlN=xgvg+xlvl=vd(xg+xl),


(7.24)


implying


xlvdvl=xgvgvd.


(7.25)


Equation (7.25) is the lever rule.29 As vdvg, xl0, and vice versa. See Exercise 7.7.


 


Note that points A and B in Fig. 7.6, which occur at separate volumes in a Pv phase diagram, are, in a PT phase diagram (such as Fig. 7.1), located at the same point on a coexistence curve. Along AB in Fig. 7.6, μl(T,P,vl)=μg(T,P,vg), and therefore from Ndμ=SdT+VdP (Gibbs-Duhem), dμ=0 (phase coexistence) and dT=0 (isotherm) implies dP=0. A horizontal, subcritical isotherm in a Pv diagram indicates liquid-gas coexistence. From the Gibbs phase rule, for a single substance in two phases, one intensive variable (the specific volume) can be varied and not disrupt phase coexistence.


The van der Waals model clearly does not have isotherms featuring horizontal segments. In 1875, Maxwell suggested a way to reconcile undulating isotherms with the requirement of phase coexistence [85]. From the Gibbs-Duhem equation applied to an isotherm ( dT=0), Ndμ=VdP, and thus for an isotherm to represent phase coexistence, we require


0=μlμg=lgdμ=lgvdP=lg[d(Pv)Pdv]=(Pv)|lglgPdv,


(7.26)


implying


Psat(T)(vgvl)=lgP(v)dv.


(7.27)


Equation (7.27) is the Maxwell equal area rule. It indicates that vl, vg, and Psat(T) should be chosen so that the area under the curve of an undulating isotherm is the same as the area under a horizontal isotherm.30 Maxwell in essence modified the van der Waals model so that it describes phase coexistence:


P(v)=kTvbav2TTc,allv;T<Tc,v>vg,v<vlPsat(T)T<Tc,vl<v<vg.


(7.28)


The Maxwell rule is an ad hoc fix that allows one to locate the coexistence region in a Pv diagram. One could object that it uses an unphysical isotherm to draw physical conclusions. It’s not a theory of phase transitions in the sense that condensation is shown to occur from the partition function.31 In an important study, Kac, Uhlenbeck, and Hemmer showed rigorously that a one-dimensional model with a weak, long-range interaction (in addition to a hard-core, short-range potential), has the van der Waals equation of state together with the Maxwell construction.[88] This work was generalized by Lebowitz and Penrose to an arbitrary number of dimensions.[89] Models featuring weak, long-range interactions are known as mean field theories (Section 7.9); the van der Waals-Maxwell model is within a class of theories, mean field theories. The energy per particle of the van der Waals gas is modified from its ideal-gas value ( 32kT) to include a contribution that’s proportional to the density, an, the hallmark of a mean-field theory; see Exercise 7.8.


 


7.3.3 Heat capacity of coexisting liquid-gas mixtures


At a point in the coexistence region (such as D in Fig. 7.6), the total internal energy of the system,


U=Nlul+Ngug,


(7.29)


where ul (ug) is the specific energy, the energy per particle uU/N, of the liquid (gas) phase at the same pressure and temperature as the coexisting phases.32 Dividing Eq. (7.29) by N,


u=xlul+xgug.


(7.30)


Compare Eq. (7.30) with Eq. (7.24), which has the same form.


To calculate the heat capacity it suffices to calculate the specific heat, cC/N. Starting from CV=U/TV,N (Eq. (1.38)), the specific heat cvCV/N=u/Tv, where v=V/N. One might think it would be a simple matter to find cv by differentiating Eq. (7.30) with respect to T, presuming that the mole fractions xl,xg are temperature independent. The mole fractions, however, vary with temperature at fixed volume. As one increases the temperature along the line associated with fixed volume vd in Fig. 7.6, xl,xg vary because the shape of the coexistence region changes with temperature; consider the lever rule, Eq. (7.25). Of course, dxl=dxg because of the constraint xl+xg=1. We have, using Eq. (7.30),


cv=(uT)v=xl(ulT)coex+xg(ugT)coex(ugul)(xlT)coex,


(7.31)


where /Tcoex indicates a derivative taken at the sides of the coexistence region.33 Evaluating these derivatives is relegated to Exercise 7.10. The final result is:


cv=xlcvlTPvlTvlTcoex2+xgcvgTPvgTvgTcoex2.


(7.32)


Equation (7.32) is a general result, not specific to the van der Waals model.



As T approaches Tc from below, TTc, the specific volumes vg, vl of coexisting gas and liquid phases tend to vc, which we can express mathematically as (vgvc)0, (vcvl)0 as (TcT)0. Can we quantify how vg,vlvc as TTc ? Because that’s something that can be measured, and calculated from theory. The critical exponent34 β characterizes how vgvc as TTc through the relation (vgvc)TTcβ. Other quantities such as the heat capacity and the compressibility show singular behavior as vg,vlvc (and the distinction between liquid and gas disappears), and each is associated with its own critical exponent, as we’ll see.


 


Define deviations from the critical point (which are also referred to as reduced variables)


tTTcTc=T¯1    ϕvvcvc=v¯1,


(7.33)


so that t0 as TTc, and ϕ0, but can be of either sign as TTc. Substituting T¯=1+t and v¯=1+ϕ into the van der Waals equation of state (7.23), we find, to third order in small quantities,


P¯=1+4t6tϕ+9tϕ232ϕ3+O(tϕ3,ϕ4).


(7.34)


We’ll use Eq. (7.34) to show that ϕ~|t| as t0 (see Eq. (7.36)).


7.4.1 Shape of the critical coexistence region, the exponent β


The Maxwell rule can be used to infer an important property of the volumes vl, vg in the vicinity of the critical point. From its form in Eq. (7.26), lgvdP=0,


0=lgv¯dP¯=lg(v¯1+1)dP¯=lgϕdP¯+lgdP¯=0lgϕ(P¯ϕ)tdϕ=ϕlϕgϕ(6t+92ϕ2)dϕ,


(7.35)


where PlPgdP=0 for coexisting phases, we’ve used Eqs. (7.33) and (7.34), and we’ve retained terms up to first order in |t| in P¯/ϕt ( ϕ~|t|). The integrand of the final integral in Eq. (7.35) is an odd function of ϕ under ϕϕ, and the simplest way to ensure the vanishing of this integral for any |t|1 is to take ϕl=ϕg, i.e., the limits of integration are symmetrically placed about ϕ=0. We can make this conclusion only sufficiently close to the critical point where higher-order terms in Eq. (7.34) can be neglected; close to the critical point the coexistence region is symmetric about vc (which is not true in general—see Fig. 7.3).35


The symmetry P(t,ϕg)=P(t,ϕg) applied to Eq. (7.34) implies 3ϕg4|t|+ϕg2=0, or


ϕg=2|t|~TTc1/2.    |t|0


(7.36)


In the van der Waals model, therefore, vgvc with the square root of TTc, and thus it predicts β=12. The measured value of β for liquid-gas critical points is β0.32 –0.34[90, 91]. We see from Fig. 7.3 that the coexistence region is well described using β=13.


7.4.2 The heat capacity exponent, α


How does the specific heat behave in the critical region? Let’s approach the critical point along the critical isochore, a line of constant volume, v=vc. As TTc, cvg,cvlcvc, and xl,xg12 (use the lever rule and the symmetry of the critical region, vcvl=vgvc). From Eq. (7.32),


cv(Tc)=cv(Tc+)Tc2limTTc[(Pvl)T(vlT)coex2+(Pvg)T(vgT)coex2],


(7.37)


where we’ve set cv(Tc+)12(cvl+cvg) as the limit TTc+, i.e., the terms in square brackets disappear for T>Tc. For the van der Waals gas, cv has the ideal-gas value for T>Tc, cv032k (see Exercise 7.8). Equation (7.37) is a general result, and is not specific to the van der Waals model.


 


For the van der Waals model, we find using Eq. (7.36),


vTcoex=vcTc1|t|,


(7.38)


where the upper (lower) sign is for gas (liquid). For the other derivative in Eq. (7.37), we can use Eq. (7.34),


PvT=PcvcP¯ϕt=Pcvc6|t|18|t|ϕ92ϕ2=12Pcvc|t|+O(|t|3/2),


(7.39)


where we’ve used Eq. (7.36). From Eq. (7.37), taking the limit,


cv(Tc)cv0=Tc12Pcvc|t|vcTc21|t|=12PcvcTc=92k,



where we’ve used Eqs. (7.38) and (7.39), and that Pcvc/Tc=38k (see Eq. (7.21)). Thus,


cvcv0=92kTTc0TTc+.


(7.40)


The specific heat is discontinuous at the van der Waals critical point—a second-order phase transition in the Ehrenfest classification scheme.


For many systems CV does not show a discontinuity at T=Tc, it diverges as TTc (from above or below),36 such as in liquid 4He at the superfluid transition,37 the so-called “λ-transition.” To cover these cases, a heat capacity critical exponent α is introduced,38


CV~TTcα.    TTc


(7.41)


The value α=0 is assigned to systems showing a discontinuity in CV at Tc, such as for the van der Waals model. The measured value of α for liquid-gas phase transitions is α0.1. Extracting an exponent from measurements can be difficult. If a thermodynamic quantity f(t) shows singular behavior as t0 in the form f(t)~Atx, the exponent is obtained from the limit


xlimt0lnf(t)lnt.


(7.42)


If x<0, f(t) diverges at the critical point; if x>0, f(t)0 as TTc.


From thermodynamics, CP>CV, Eq. (1.42). The heat capacity CV for the van der Waals gas is the same as for the ideal gas (Exercise 7.8). CP, however, has considerable structure for T>Tc; see Eq. (P7.4). In terms of reduced variables, Eq. (P7.4) is equivalent to (show this):


CPCVNk=(1+ϕ)3(1+t)34ϕ2+ϕ3+t(1+ϕ)3.


(7.43)


Because ϕ0 as |t|, Eq. (7.36), we see that CP diverges as TTc+, with


CP~TTc1.    (TTc+)


(7.44)


Equation (7.44) is valid for T>Tc; CP isn’t well defined in the coexistence region.39


 


7.4.3 The compressibility exponent, γ


We’ve introduced two exponents that characterize critical behavior: α for the heat capacity and β for vgvl. Is there a γ? One can show that at the van der Waals critical point,


Pv|T=Tc,v=vc=2Pv2|T=Tc,v=vc=0.


(7.45)


The critical isotherm therefore has an inflection point at v=vc, as we see in Figs. 7.4 or 7.6. The two conditions in Eq. (7.45) define critical points in fluid systems. One can show that


(Pv)t=2a27b3[34ϕ2+ϕ3+t(1+ϕ)3(1+ϕ)3(1+32ϕ)2].


(7.46)


Thus, (P/v)t~t as t0. Its inverse, the compressibility βT=(1/v)(v/P)T therefore diverges at the critical point,40which is characterized by the exponent γ,


βT~TTcγ.


(7.47)


The van der Waals model predicts γ=1; experimentally its value is closer to γ=1.25.


7.4.4 The critical isotherm exponent, δ


The equation of state on the critical isotherm is obtained by setting t=0 in Eq. (7.34), P¯132ϕ3, implying


P¯1=PPcPc=32vvcvc3~vvcvcδ,


(7.48)


where δ is the conventional symbol for the critical isotherm exponent. The van der Waals model predicts δ=3, whereas for real fluids [91], δ4.7 –5.


The definition of the critical exponents α,β,γ,δ and their values are summarized in Table 7.1.






























Table 7.1 Critical exponents α,β,γ,δ for liquid-vapor phase transitions
Exponent   Definition Van der Waals model Measured
α CV~TTcα α=0 α0.1
β vgvl~TTcβ β=12 β0.33
γ βT~TTcγ γ=1 γ1.25
δ PPc~VVcδ δ=3 δ4.75


We studied paramagnetism in Section 5.2, where independent magnetic moments interact with an external magnetic field. The hallmark of paramagnetism is that the magnetization M vanishes as the applied field is turned off, at any temperature. We saw in the one-dimensional Ising model that adding interactions between spins does not change the paramagnetic nature of that system, Eq. (6.84). Ferromagnets display spontaneous magnetization—a phase transition—in zero applied field, from a state of M=0 for temperatures T>Tc, to one of M0 for T<Tc. The two-dimensional Ising model shows spontaneous magnetization (see Section 7.10), so that dimensionality is a relevant factor affecting phase transitions. Indeed, we’ll show that one-dimensional systems cannot support phase transitions (Section 7.12). The two-dimensional Ising model has a deserved reputation for mathematical difficulty. Is there a simple model of spontaneous magnetization?


 


In 1907, P. Weiss introduced a model that bears his name, the Weiss molecular field theory. Figure 7.7 shows part of a lattice of magnetic moments. Weiss argued that if a system is magnetized, the aligned dipole moments of the system would produce an internal magnetic field, Bmol, the molecular field,41 that’s proportional to the magnetization, Bmol=λM, where the proportionality constant λ will be inferred from the theory. The internal field Bmol would add to the external field, B. Reaching for Eq. (5.18) (the magnetization of independent magnetic moments),


M=Nμz=NμLβμ(B+λM),


(7.49)

Figure 7.7

Figure 7.7Molecular field Bmol is produced by the surrounding magnetic dipoles of the system.


where L(x) is the Langevin function. Equation (7.49) is a nonlinear equation of state, one that’s typical of self-consistent fields where the system responds to the same field that it generates.42


Because we’re interested in spontaneous magnetization, set B=0 in Eq. (7.49):


M=NμL(βμλM).


(7.50)


Does Eq. (7.50) possess solutions? It has the trivial solution M=0 ( L(0)=0). If Eq. (7.50) is to represent a continuous phase transition, we expect that it has solutions for M0. Using the small-argument form of L(x)13x145x3 (see Eq. (P5.1)), Eq. (7.50) has the approximate form for small M,


MaMbM3    M1a+bM2=0,


(7.51)


where aβNμ2λ/3 and bβ3μ4Nλ3/45 are positive if λ>0. If a1, Eq. (7.51) has the trivial solution. If, however, a>1, it has the nontrivial solution M2(a1)/b. Clearly a=1 represents the critical temperature. Let43


kTc13Nμ2λ    λ=3kTcNμ2.


(7.52)


Combining Eqs. (7.52) and (7.50), and letting yM/(Nμ), we have the dimensionless equation:


y=L3TcTy.


(7.53)


Figure 7.8 shows a graphical solution of Eq. (7.53). For TTc, there is only the trivial solution, whereas for T<Tc, there is a solution with M0 for B=0. Spontaneous magnetization thus occurs in the Weiss model. Just as with liquid-vapor phase transitions, exponents characterize the critical behavior of magnetic systems. We now consider the critical exponents of the Weiss model.

Figure 7.8

Figure 7.8Graphical solution of Eq. (7.53). Spontaneous magnetization occurs for T<Tc.


 


7.5.1 The magnetization exponent, β


To find β is easy—it’s implied by Eq. (7.51) for T<Tc. In the limit TTc,


M2=(53N2μ2)(TcTTc)~(TcTTc)2β,(TTc)


(7.54)


implying β=12 for the Weiss model. Experimental values of β for magnetic systems are in the range 0.30–0.325[94, 95].


7.5.2 The critical isotherm exponent, δ


At T=Tc, M0 as B0; the vanishing of M in characterized by the critical exponent δ, defined as B~Mδ at T=Tc. To extract the exponent, we need to “dig out” B from the equation of state, Eq. (7.49), for small M. Combining Eq. (7.52) with Eq. (7.49), where y=M/(Nμ),


y=LβμB+3TcTy.


(7.55)


We can isolate the field term B in Eq. (7.55) by invoking the inverse Langevin function L1(y):


βμB=L1(y)3TcTy.


(7.56)


A power series representation of L1 can be derived from the Lagrange inversion theorem [50, p14], where for 1<y<1,


L1(y)=k=1ykk!dk1dxk1xL(x)kx=0k=1ckyk=3y+95y3+297175y5+O(y7).


(7.57)


Additional coefficients ck are tabulated in [96]. Combining the small-y form of L1 from Eq. (7.57), with Eq. (7.56), we find


βμB=3t1+ty+95y3+O(y5),


(7.58)


where t is defined in Eq. (7.33). On the critical isotherm t=0, and thus from Eq. (7.58),


B=(9kTc5μ)(MNμ)3.(|M|Nμ)


(7.59)


Equation (7.59) should be compared with Eq. (7.48), the analogous result for fluids. The Weiss model predicts δ=3, whereas experimentally δ4 –5.


7.5.3 The susceptibility exponent, γ


The magnetic susceptibility χ=(M/B)T, see Eq. (1.50), is the magnetic analog of the compressibility of fluids, Eq. (1.33). For small M, we have by implicitly differentiating Eq. (7.58),


βμ=t1+t1Nμχ+OM2χ.



We defined χ in Eq. (1.50) as the isothermal susceptibility, but the value of B was unrestricted. What’s usually referred to as the magnetic susceptibility is the zero-field susceptibility. As TTc,


χ=MBT,B=0=Nμ23k1TTcCTTc~TTcγ,


(7.60)


where C=Nμ2/(3k) is the Curie constant44 for this system (see Section 5.2). Equation (7.60) is the Curie-Weiss law;45 it generalizes the Curie law, Eq. (5.15), M=CH/T. The Weiss model predicts γ=1.


7.5.4 The heat capacity exponent, α


From the Helmholtz energy F=UTS and the magnetic Gibbs energy Gm=FBM (see Section 1.9), we have the identities (using dU=TdS+BdM)


S=FTM  B=FMTS=GmTB  M=GmBT.


(7.61)


Combining Eqs. (P1.13) and (P1.14) with the results in Eq. (7.61),


CM=TSTM=T2FT2MCB=TSTB=T2GmT2B.


(7.62)


Which formula46 should be used to calculate the heat capacity in the Weiss model?


Equation (7.58) is a power series for B(M), which when combined with B=F/MT (Eq. (7.61)) provides an expression that can be integrated term by term:


B=FMT=1βμ3t1+ty+9kT5μy3+O(y5)=3kNμ2(TTc)M+9kT5μ1(Nμ)3M3+O(M5),



implying,


F(T,M)=F0(T)+3k2Nμ2(TTc)M2+9kT20μ1(Nμ)3M4+O(M6),


(7.63)


where F0(T) is the free energy of the non-magnetic contributions to the equilibrium state of the system “hosting” the magnetic moments (and which cannot be calculated without further information about the system).


 


For TTc, M=0, and for TTc, M2 is given by Eq. (7.54), so that, from Eq. (7.63),


F(T)={F0(T)TTcF0(T)52NkTct2TTc.


(7.64)


Using Eq. (7.62),


CMC0=0TTc5NkTTc.


(7.65)


where C0 is the heat capacity of the non-magnetic degrees of freedom. Thus, there is a discontinuity in CM for the Weiss model.


The critical exponent α for magnetic systems is, by definition, associated with CB (not CM) as B0,


CB=0~TTcα.


(7.66)


One might wonder why CB is used to define the exponent α, given the correspondence with fluid systems VM, PB. For TTc, B=0 implies M=0, and for T<Tc the “two phase” M=0 heat capacity also corresponds to B=0 because with ferromagnets there is up-down symmetry. For fluids, we considered the heat capacity of coexisting phases, each of which has a different heat capacity; see Eq. (7.32). For the Weiss model, we can use Eq. (1.51) to conclude that CB also has a discontinuity at the critical point. Thus, α=0 for the Weiss model.


The critical exponents of the Weiss model have the same values as those in the van der Waals model. Magnets and fluids are different physical systems, yet they have the same critical behavior. Understanding why the critical exponents of different systems can be the same is the central issue of modern theories of critical phenomena; see Chapter 8. Before proceeding, we introduce two more exponents that are associated with correlations in the critical region.47


Only gold members can continue reading. Log In or Register to continue

Jul 18, 2021 | Posted by in General Engineer | Comments Off on 7 ▪ Phase transitions and critical phenomena
Premium Wordpress Themes by UFO Themes