Phase Transitions. µ a (P c (T ), T ) µ b (P c (T ), T ), (3) µ a (P, T c (P )) µ b (P, T c (P )). (4)

Similar documents
Grand Canonical Formalism

Physics 408 Final Exam

f(t,h) = t 2 g f (h/t ), (3.2)

Identical Particles. Bosons and Fermions

III. The Scaling Hypothesis

The Clausius-Clapeyron and the Kelvin Equations

Classical Thermodynamics. Dr. Massimo Mella School of Chemistry Cardiff University

Chapter 6. Phase transitions. 6.1 Concept of phase

Removing the mystery of entropy and thermodynamics. Part 3

Overview of phase transition and critical phenomena

Thermodynamic condition for equilibrium between two phases a and b is G a = G b, so that during an equilibrium phase change, G ab = G a G b = 0.

The (magnetic) Helmholtz free energy has proper variables T and B. In differential form. and the entropy and magnetisation are thus given by

CHEM-UA 652: Thermodynamics and Kinetics

Chapter 6 Thermodynamic Properties of Fluids

Physics 127b: Statistical Mechanics. Renormalization Group: 1d Ising Model. Perturbation expansion

Phase Transitions. Phys112 (S2012) 8 Phase Transitions 1

4.1 Constant (T, V, n) Experiments: The Helmholtz Free Energy

Physics 212: Statistical mechanics II Lecture XI

1. Heterogeneous Systems and Chemical Equilibrium

(# = %(& )(* +,(- Closed system, well-defined energy (or e.g. E± E/2): Microcanonical ensemble

CHAPTER 4 Physical Transformations of Pure Substances.

CHEM-UA 652: Thermodynamics and Kinetics

Physics Nov Phase Transitions

Phase Equilibrium: Preliminaries

Chemistry 2000 Fall 2017 Test 2 Version B Solutions

At this point, we've developed the tools and basic concepts necessary to apply

Math 123, Week 9: Separable, First-Order Linear, and Substitution Methods. Section 1: Separable DEs

UNDERSTANDING BOLTZMANN S ANALYSIS VIA. Contents SOLVABLE MODELS

Physics 127a: Class Notes

Thus, the volume element remains the same as required. With this transformation, the amiltonian becomes = p i m i + U(r 1 ; :::; r N ) = and the canon

Remember next exam is 1 week from Friday. This week will be last quiz before exam.

For an incompressible β and k = 0, Equations (6.28) and (6.29) become:

1. Thermodynamics 1.1. A macroscopic view of matter

Physics 53. Thermal Physics 1. Statistics are like a bikini. What they reveal is suggestive; what they conceal is vital.

Introduction Statistical Thermodynamics. Monday, January 6, 14

Physics 127b: Statistical Mechanics. Second Order Phase Transitions. The Ising Ferromagnet

The Second Law of Thermodynamics (Chapter 4)

Liquids. properties & structure

Chapter 5 - Systems under pressure 62

Introduction. Chapter The Purpose of Statistical Mechanics

University of Illinois at Chicago Department of Physics. Thermodynamics and Statistical Mechanics Qualifying Examination

Thermodynamics of phase transitions

The Ising model Summary of L12

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms.

Lecture Phase transformations. Fys2160,

Entropy and the second law of thermodynamics

University of Illinois at Chicago Department of Physics SOLUTIONS. Thermodynamics and Statistical Mechanics Qualifying Examination

Homework 8 Solutions Problem 1: Kittel 10-4 (a) The partition function of a single oscillator that can move in three dimensions is given by:

(i) T, p, N Gibbs free energy G (ii) T, p, µ no thermodynamic potential, since T, p, µ are not independent of each other (iii) S, p, N Enthalpy H

Physics 119A Final Examination

Phase Transitions: A Challenge for Reductionism?

THERMODINAMICS. Tóth Mónika


Reaction Dynamics (2) Can we predict the rate of reactions?

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space.

1 Phase Transitions. 1.1 Introductory Phenomenology

Step 1. Step 2. g l = g v. dg = 0 We have shown that over a plane surface of water. g v g l = ρ v R v T ln e/e sat. this can be rewritten

THE SECOND LAW OF THERMODYNAMICS. Professor Benjamin G. Levine CEM 182H Lecture 5

Phase Change (State Change): A change in physical form but not the chemical identity of a substance.

Section 3 Entropy and Classical Thermodynamics

Separation of Variables in Linear PDE: One-Dimensional Problems

Homework Hint. Last Time

MME 2010 METALLURGICAL THERMODYNAMICS II. Fundamentals of Thermodynamics for Systems of Constant Composition

A Phase Transition in Ammonium Chloride. CHM 335 TA: David Robinson Office Hour: Wednesday, 11 am

Temperature and Heat. Prof. Yury Kolomensky Apr 20, 2007

Thermodynamics and Statistical Physics. Preliminary Ph.D. Qualifying Exam. Summer 2009

CHEMISTRY 443, Fall, 2014 (14F) Section Number: 10 Examination 2, November 5, 2014

Thermodynamic Functions at Isobaric Process of van der Waals Gases

Thermodynamic Properties

Chapter 4: Going from microcanonical to canonical ensemble, from energy to temperature.

ChE 503 A. Z. Panagiotopoulos 1

Physics 360 Review 3

CHAPTER V. Brownian motion. V.1 Langevin dynamics

to satisfy the large number approximations, W W sys can be small.

Renormalization Group for the Two-Dimensional Ising Model


21 Lecture 21: Ideal quantum gases II

CHAPTER 11: Spontaneous Change and Equilibrium

Physics 408 Final Exam

Thermodynamics System Surrounding Boundary State, Property Process Quasi Actual Equilibrium English

Chemistry 593: The Semi-Classical Limit David Ronis McGill University

On the local and nonlocal components of solvation thermodynamics and their relation to solvation shell models

Thermodynamics (Classical) for Biological Systems Prof. G. K. Suraishkumar Department of Biotechnology Indian Institute of Technology Madras

Chemical Equilibria. Chapter Extent of Reaction

Basic Thermodynamics Module 1

Table of Contents [ttc]

MACROSCOPIC VARIABLES, THERMAL EQUILIBRIUM. Contents AND BOLTZMANN ENTROPY. 1 Macroscopic Variables 3. 2 Local quantities and Hydrodynamics fields 4

Fundamental equations of relativistic fluid dynamics

PHY101: Major Concepts in Physics I

A.1 Homogeneity of the fundamental relation

Chemistry 2000 Fall 2017 Test 2 Version A Solutions

Caltech Ph106 Fall 2001

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 30 Jun 2003

Theoretical Statistical Physics

Chapter 15. Landau-Ginzburg theory The Landau model

9.1 System in contact with a heat reservoir

8.044 Lecture Notes Chapter 8: Chemical Potential

Basic Thermodynamics Prof. S. K. Som Department of Mechanical Engineering Indian Institute of Technology, Kharagpur. Lecture No 16

ChE 210B: Advanced Topics in Equilibrium Statistical Mechanics

Collective behavior, from particles to fields

Transcription:

Phase Transitions A homogeneous equilibrium state of matter is the most natural one, given the fact that the interparticle interactions are translationally invariant. Nevertheless there is no contradiction with the fundamentals of the Statistical Mechanics if under some conditions we have at a fixed volume different phases of one and the same substance: liquid and vapor, solid and liquid, liquid and vapor, etc. What are the conditions for such a situation to occur, and why do different phases exist in principle? First, let us take the existence of different phases for granted, and consider the conditions for their coexistence. Each phase is assumed to be homogeneous and characterized by its equation of state in the form µ = µ(p, T ), (1) in which all the three variables are intensive. Suppose we have two coexisting phases, a and b. Then, each of them, with respect to the other one, simultaneously plays the following three roles: (i) a piston, (ii) a heat bath, and (iii) a particle reservoir. This means that (i) the pressures, (ii) the temperatures, and (iii) the chemical potentials of the two phases coincide, leading to the following condition of the phase coexistence in terms of the equations of state: µ a (P, T ) = µ b (P, T ). (2) Eq. (2) implies that for a given temperature, T, there can exist only one special critical pressure, P c (T ), at which the coexistence of the two phases is possible. Equivalently, for a given pressure, P, there can exist only one critical temperature, T c (P ). The functions P c (T ) and T c (P ) can be found from (2) in an implicit form: µ a (P c (T ), T ) µ b (P c (T ), T ), (3) µ a (P, T c (P )) µ b (P, T c (P )). (4) Hence, the relation (2) defines a critical line in the (P, T )-plane on which and only on which the coexistence of two phases is possible. Clearly, coexistence of three different phases is possible only at an isolated point so-called triple point on the (P, T )-plane, as in this case we have one more independent constraint, say, µ b (P, T ) = µ c (P, T ). (5) Eqs. (2) and (5) can be simultaneously satisfied only at an isolated point: two independent equations for two unknown variables. In a general case, the coexistence of more than three different phases is impossible because we will get three or more independent equations for just two variables. We see that at a fixed pressure there is only one special temperature, T c (P ), at which two (or three) phases can coexist. Hence, if we pass this special temperature the system will change its phase in a jump-like way. This phenomenon is known as the first-order phase transition. Now we are in a position to understand the reason for a phase transition to occur, and, in particular, the reason for existing different phases, since there is no phase transitions without phases. Fixed pressure and varying temperature are very convenient variables, because the state of the system remains homogeneous up to a single special point T c. The reason for different phases to exist is thermodynamic favorability. In the space of states of a macroscopic system there can be two or more different regions corresponding to qualitatively different properties of the system (liquid, vapor, solid, ferromagnetic/paramagnetic, etc.). In accordance with Gibbs distribution, the system can be found in each of these regions with the probability proportional 1

to the contribution to the total partition function, associated with the corresponding region. Given two phases, a and b, we thus write the corresponding probabilities as P a = Z a Z a + Z b, P b = Z b Z a + Z b. (6) Here we face a technical problem: We know how to calculate the partition function at a fixed volume, as the volume is a natural parameter fixing the spectrum of the energy eigenvalues. But now we want to fix pressure, not volume. A trick is to treat the pressure as resulting from a constant force (equal to pressure times piston area) applied to our system. We thus get a situation of an external potential equal to force coordinate = P surface area coordinate = P V. (7) That is for a system of the volume V we get the extra contribution of P V to the energy (and Helmholtz free energy), or, equivalently, the factor of exp(p V/T ) to the partition function. Recalling that F + P V = G, (8) where G is the Gibbs free energy, we conclude that at a fixed pressure the effective partition function for the phase a reads Z a = e Ga/T, (9) and the same for the phase b. Hence, P a P b = Z a Z b = e (G b G a)/t. (10) The probabilities become equal at the point where G a (P, T ) = G b (P, T ), (11) and are radically different away from this point, since the difference G b G a is an extensive quantity, proportional to the total number of particles. If it is non-zero, then it is macroscopically large. Hence, at the critical point the systems jumps from one phase into the other one. In accordance with Eq. (10), the most favorable phase is the phase with the lowest Gibbs free energy. Is there any connection between the condition (11) and the condition (2) of phase coexistence? Yes, and quite a trivial one, because G µn. (12) Problem 47. Establish the identity (12). Rewriting (10) as P a P b = Z a Z b = e N(µ b µ a)/t, (13) we see that at the fixed pressure the most favorable phase is the phase with a smaller chemical potential. The Gibbs free energies of the two phases are equal at the phase transition point. What about the rest of quantities? Taking into account that G = E T S + P V, (14) and that, by definition, the pressure and temperature remain the same at the phase transition point, from (11) we get E a T S a + P V a = E b T S b + P V b. (15) 2

Note that neither energy, nor entropy, nor volume is supposed to be the same in the two phases. Eq. (16) can be rewritten as E b E a = T (S b S a ) + P (V a V b ). (16) In this form it describes the energy balance. The left-hand side is the change of energy. The second term in the right-hand side is the work performed over the system. Correspondingly the term L ab = T (S b S a ) (17) is the heat transferred to the system from the heat bath during the transition. It is called the latent heat. During the I-order phase transition the heat transfer does not result in any temperature change. This means that the heat capacity behaves like a δ-function of (T T c ). Problem 48. Derive the Clapeyron equation: Hint. Differentiate the identity (3), remembering (12). dp c dt = L ab T (V b V a ). (18) Consider some point on the phase transition line T = T c (P ). At the given temperature and pressure the system can have two different volumes, V a, and V b. What happens if starting from, say, volume V a < V b, we will be keeping temperature constant while increasing volume, V? The system cannot be totally in the phase b, until V reaches the value V b. It cannot be totally in the phase a either, since V > V a. The only possibility is the phase separation, when a part of the system volume corresponds to the phase b, while the rest of the volume is in the phase a. And this is precisely the situation of the phase coexistence, described by the relation (2). From (2) it follows that if T is fixed, then the pressures of the two coexisting phases correspond to the critical pressure P c (T ) and thus are independent of volume. The same is true for the densities, since the temperature and pressure unambiguously define the density of each of the two phases. Hence, the only quantities that are sensitive to V in the region V [V a, V b ] are the volumes V 1 and V 2 of the components a and b, respectively. These volumes are easily found from the relation V 1 V a + V 2 V b = 1, (19) and an obvious equation V 1 + V 2 = V. (20) Problem 49. Show that Eq. (19) follows from the requirement that the total number of particles be fixed. The Critical Point. Second-Order Phase Transitions Suppose we take a fixed volume with coexisting liquid and vapor, and start to increase the temperature. At high enough T, there should be no phase separation: in the limit of T the system behaves like an ideal gas, because the kinetic energy of particles is much larger than the potential energy of their interaction. Hence, with increasing temperature we always cross a phase transition line. There are two generic scenarios of what will be happening: (i) the liquid component expands until the vapor component vanishes; (ii) the liquid component evaporates and vanishes. By continuity argument, we conclude that for the given number of particles, N, there should exist such a critical volume, V (N) it is a bit more physical to talk of a critical density, n, because V N/n, that the volumes of both coexisting phases remain finite up to the very phase transition point. The phase transition point at n = n turns out to be different from the generic (first-order) liquid-vapor phase transition point. At this point there is no latent heat. All the thermodynamic potentials remain continuous. When approaching this point, the densities of liquid and vapor approach each other, so that beyond this point it becomes fundamentally impossible to tell one from another. 3

The critical point in the liquid-vapor system corresponds to what is known as the second-order phase transition. A generic feature of the second-order phase transitions, as opposed to the first-order transitions, is the absence of jumps in thermodynamic potentials and densities. Instead of jumps there take place anomalous fluctuations with fractal structure. In terms of liquid and vapor, the fractal structure of fluctuations looks as follows. In the liquid domain of the phase separated system there occur and disappear large regions of vapor. In these regions of vapor there are somewhat smaller regions of liquid, and so on, down to the microscopic scale. Staring from the interatomic distances, the fractal structure of fluctuations ranges to some of distance called correlation radius, r c. At distances smaller than r c it is simply impossible to tell liquid from vapor, and, correspondingly, whether the temperature is below or above the critical point! Correlation radius is a function of the vicinity to the critical point, which blows up at the critical point. For example, with temperature approaching T at fixed n = n, the correlation radius behaves like r c 1 T T ν, (21) where the critical exponent ν is one and the same for both T > T and T < T regions, while the proportionality coefficients are different. Another generic feature of the second-order phase transitions is a non-analytic behavior (often divergence) of the heat capacity: 1 C T T α. (22) The difference between the density of vapor (liquid) and the critical density behaves like n liq n = n n vap (T T ) β, (23) For the liquid vapor critical point: ν 2/3, α 0.1, β 0.3. The relation (23) completely describes the behavior of the system in the phase separation region in the vicinity of the critical point. Indeed, suppose the volume of the system is a bit different from the V, which is equivalent to saying that the total number of particles is a bit different from its critical value N. We are in a phase separation region, so we have two phases with the volumes V 1 and V 2 for the liquid and vapor, respectively. We have a straightforward relation of the particle balance: n liq V 1 + n vap V 2 = N. (24) An important fact about this relation see previous discussion is that n liq and n vap depend only on temperature, and not on V 1 and V 2. This means that we can use (23) for n liq and n vap even if we are not exactly at the critical volume [but still rather close to it, since this relation works only when T T T ]. So, identically rewriting (24) as (n liq n )V 1 + (n vap n )V 2 = N N, (25) using (23), and eliminating V 2 by V 1 + V 2 = V, we find the relation V 1 V 1 2 which yields the fraction of the volume occupied by the liquid at given (T, n, V ). n n (T T ) β (26) Problem 50. Show that Eq. (23) implies that the boundary of the phase coexistence region in the (V, T ) plane obeys in the vicinity of the point (V, T ) the equation: T c T V c V 1/β. (27) Here T c = T c (V c ) is the equation for the boundary [which can be equivalently read as V c = V c (T c )]. Hint 1. Note that boundary actually consists of two separate curves meeting each at the critical point. One of the curves corresponds to V 1 = V (and thus V 2 = 0) and another one is the curve where V 2 = V (and thus V 1 = 0). Hint 2. Do not forget to take into account that all our relations are valid only when T T T, V V V, 4

so that everywhere except for the differences T T and V V one can replace T with T and V with V. Scale Invariance. Hyperscaling Relation Consider a sequence of linear system sizes, {L n }, (n = 0, 1, 2,...), that differ by some given factor, say, 2 for definiteness. That is L n = 2 n l 0 (with l 0 being some large as compared to interatomic distances, but still microscopic scale). Let Z n be the partition function for the system size L n. For a normal macroscopic system we would have Z n+1 = (Z n ) 2d (normal), (28) where d is the spatial dimension. [Note that in d dimensions the system of the linear size 2L consists of 2 d subsystems of the size L.] Eq. (28) is basically the definition of a macroscopic system as an ensemble of independent subsystems. At the critical point of a second-order phase transition we can formally write Z n+1 = (Z n ) 2d ξ (critical), (29) where the dimensionless quantity ξ is a correction to Eq. (28) coming from the fact that there are macroscopically large fluctuations. A crucial assumption about ξ, that holds true for all standard second-order phase transitions is that ξ is one and the same for all n s. This assumption follows from a bit more general assumption of scale invariance of the fluctuations at the second-order critical point. Scale invariance here means independence of the length scale. For the free energies F n = T ln Z n we then have with F n+1 = af n + f, (30) a = 2 d, f = T ln ξ. (31) The recursive relation (30) allows one to explicitly express F n in terms of F 0 = T ln Z 0 and f: F n = a n F 0 + an 1 a 1 f. (32) Now if we are in a close vicinity of the critical point, but not exactly at the critical point, then expressions (29)-(32) work only up to n = n corresponding to the correlations radius: n = log 2 (r c /l 0 ). (33) For L > r c normal scaling takes place, with F directly proportional to the system volume. Hence, the free energy F corresponding to a system size L r c can be written as In accordance with (32)-(33) this means F = F (L) = (L/r c ) d F n. (34) ( F 0 + f a 1 ) ( L l 0 ) d f a 1 ( ) d L. (35) The first term in the r.h.s. of Eq. (35) does not contain r c and thus is not sensitive to approaching the criticality. The second term, which we denote as F sing, contains r c and thus demonstrates a certain singularity at the critical point. It is convenient to characterize the vicinity to the critical point T c by the dimensionless parameter Then, in accordance with (21) we have r c 1/τ ν and thus r c τ = T T c T c. (36) F sing τ dν. (37) 5

Now we readily find the singular parts of entropy and heat capacity S sing F sing τ τ dν 1, (38) C sing S sing τ dν 2. (39) τ Comparing Eq. (39) to the definition (22) of the critical exponent α, we arrive at the celebrated hyperscaling relation α = 2 dν (40) which establishes a dependence between the critical exponents ν and α. 6